Conversion of Isobutylene to Octane-Booster Compounds after Methyl

Sep 26, 2016 - ... feedstock is used for the production of another octane booster compound such as isooctane with a more environmental-friendly reputa...
25 downloads 0 Views 1MB Size
Subscriber access provided by CORNELL UNIVERSITY LIBRARY

Review

Conversion of isobutylene to octane-booster compounds after MTBE phaseout: The role of heterogeneous catalysis Hilman Ibnu Mahdi, and Oki Muraza Ind. Eng. Chem. Res., Just Accepted Manuscript • DOI: 10.1021/acs.iecr.6b02533 • Publication Date (Web): 26 Sep 2016 Downloaded from http://pubs.acs.org on October 8, 2016

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Industrial & Engineering Chemistry Research is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Conversion of isobutylene to octane-booster compounds after MTBE phaseout: The role of heterogeneous catalysis Hilman Ibnu Mahdi1, Oki Muraza1* 1

Chemical Engineering Department and Center of Research Excellence in Nanotechnology, King Fahd University of Petroleum & Minerals, Dhahran 30261, Saudi Arabia

Abstract

Over the last decades, the global demand of the production of MTBE was significantly decreased from 30 million tons per year (MTPY) in 1995 to 12 million tons per year (MTPY) in 2011 because of the MTBE phaseout in the United States and other countries. The MTBE was banned in some countries because it produces health and environmental pollution and contaminates water and air quality. After MTBE phaseout, isobutylene feedstock is used for the production another octane booster compound such as isooctane with more environmental friendly reputation. Isooctane is obtained over dimerization, trimerization, and oligomerization of isobutylene followed by a hydrogenation step. Dimerization of isobutylene to isooctane is a simple and low-cost effective technology using heterogeneous catalysts such as zeolites, metal oxides, and resins. Zeolites are potential catalyst with high selectivity, high stability, easy regeneration, and low-cost production. The conversion and selectivity can be increased with higher acid strength, larger dimensional channel, and higher pore size. USY zeolite (FAU topology) is a promising zeolite framework to be improved in dimerization of isobutylene. Further improvement on USY zeolite has been emphasized to develop better dimerization catalysts. Keywords: Isooctane; zeolite; isobutylene; heterogeneous catalysis. *Corresponding author. Tel.:+966 13 860 7612; e-mail address: [email protected] 1

1.

Introduction 1.1 Conventional MTBE production for gasoline additive One of the most important applications of isobutylene is its usage as an octane booster

compound in the production of methyl tert-butyl ether (MTBE) 1. MTBE is produced from isobutylene and methanol using various catalysts number of gasoline

10

2-9

and further used to increase the octane

. The global capacities for MTBE production, which was started in Italy,

were increased from 50,000 tons/year

10

to 100,000 tons/year

11

. Its production significantly

increased to approximately 12 million tons per year (MTPY) in 1992, 20 MTPY in 1993 – 1994 11-13

, and 30 MTPY in 1995, where 53% was originated from USA capacity 10. However, in the last decade, the worlwide demand for MTBE production significantly

decreased from approximately 22 MTPY in 2000, 18 MTPY in 2003, and 17 MTPY in 2004 14 to 12 MTPY in 2011 15 because of the MTBE ban in the US and Canada 6, 14 as presented on Figure 1. MTBE demand is predicted to grow up to 2020 with average annual growing rate (AAGR) about 5.9% per year. Approximately, 62% of the worldwide demand for MTBE production is in the Asia Pacific region especially China 15. MTBE is commercially produced using variety of catalysts such as Lewatit K-2631 7, CVT resin, Amberlyst 15 (A-15)

16, 17

heteropolyacid (Dawson acid)

20

, Bowex 50w, Lewatit SPC 118

11, 18

, montmorillonites

19

and

based catalysts, clays treated using Al3+, Fe3+, Cr3+, Cu2+, Pb2+,

Ni2+, Co2+, Ca2+, and Na+ cations 5, fluorine-promoted SiO2-Al2O3 catalyst

21

, and sulfur

promoted ZrO2 catalyst 21. Development of MTBE production was also carried out using zeolites such as Y, ZSM-5, mordenite, Beta, NaA, and ZSM-11

22-25

to replace ion exchange resins as alternative catalysts

because of their superior performance

11

and highly selective catalysts even at low

2

isobutylene/methanol ratio

22

. The ZSM-5 and ZSM-11 are better than A-15 resin because of

better thermal stability, higher selectivity, and low sensitivity to methanol:isobutylene molar ratio

23

. Other zeolites such as boron pentasil zeolite and Beta zeolite with Si/Al ratio of 15.8

were applied in MTBE synthesis via in situ MAS NMR spectroscopy 26-28 and 13 X zeolite was utilized as an adsorbent in oxygenate removal for the production of MTBE over adsorptive separation 29. 1.2 Availability of feedstocks in MTBE plant The worldwide demand for isobutylene production has been dramatically increased from 10 MTPY

30, 31

to 30 MTPY

32

because it was widely utilized for many processes in the chemical

industry such as 41% for MTBE or ETBE, 47% for alkylate gasoline, 9% for synthetic rubber, and 3% for other products

30, 32-34

. Poly-isobutene, isoprene, butyl rubber, tert-butyl mercaptans,

tert-butanol, antioxidants, and isooctane were also produced from isobutylene 30, 35-40. At least, 650,000 barrels of isobutylene were produced in the USA in 2011 35, 41. Isobutylene was obtained over isomerization and catalytic dehydrogenation using butanes

42

with different

catalysts such as chromia alumina and zeolite membrane reactors (ZMR) treated by Pd/Ag Isobutylene was also mostly produced over skeletal isomerization of n-butenes

1, 44, 45

38, 43

.

over

various catalysts namely WOx/ γ-Al2O3 46, WOx/ ferrierite 47, MCM-22 48, 49, ZSM-22 50, Omega, SAPO 11, SAPO 34, ZSM-5 zeolite 51, and natural zeolite (clinoptilolite) 52-54. The performance of natural clinoptilolite zeolite can be improved over 1 N HCl in the synthesis of the dealuminated natural clinoptilolite zeolite at 120 oC for 1.5 h

55

. Other processes such as

dehydroisomerization of butanes over ZSM-5/Pt 56, H-TON/(Pt and Cu) 57, and H–Y/(Pt and Sn) 58

and direct conversion of n-butane over MCM-22/Pt 59 also produced isobutylene.

3

Similar with isobutylene, the production for methanol as a feedstock was also high

60, 61

.

Historically, the worldwide production of methanol was dramatically increased from 22.1 MTPY in 1991 60 to 47 MTPY in 2011 62 and it was utilized for feedstock of chemical reaction of 70%, 25% for MTBE, and 9% for acetic acid

61, 63

. The highest demand of methanol in the USA

happened in 1940s with more than 180 MTPY 64. Apart from its utilization as a feedstock in MTBE plant, methanol is also a promising automotive fuel in the future because of its unique physicochemical properties

65

.

Approximately, 85% of methanol was mixed with 15% of unleaded gasoline to the use in vehicle fuel

61, 65

. Methanol can be obtained from various sources namely natural gas, coal, petroleum,

and biomass over different catalysts such as Al2O3/Ni, Cu chromite, and ZnO/Cu

60, 66-71

.

Approximately, more than 75% methanol was currently produced over natural gas process 65, 72. The current demand of renewable sources for isobutylene production has been considered in sustainable production because of high pollution and limited fossil resources. One of the renewable sources applied to obtain a green isobutylene is isobutylene production via fermentation and using microorganism as reported by previous researchers 30, 73-80. Those reports show that isobutylene production could be obtained in the fermentative process where waste from industry was used as nutrient sources in the cultivation 76, 81.

1.3 Other prospective octane booster compounds in gasoline after MTBE phaseout MTBE was banned in the USA and Canada during the last decade because it results health and environmental pollution and leads to contamination in underground water quality because of nitrogen oxide, carbon monoxide, and hydrocarbons

13, 86

80, 82-85

and air

. Because of this

reason, the worldwide MTBE growth will be declined in the future. Hence, other potential

4

processes that can be developed from derivatives of isobutylene, other than MTBE, as the octane booster are considered. Ethyl tert-butyl ether (ETBE) and isooctane can be utilized as an octane booster to replace MTBE because these products are the more environmental friendly. ETBE is obtained from isobutylene synthesis with ethanol and isooctane is obtained from isobutylene synthesis with isobutane. Sensitivity analysis of ETBE synthesis can be modeled with the assistance of Aspen PLUS 87. The ETBE production has been studied by many researchers using various catalysts. Zeolites and resins utilized in MTBE synthesis were also used in the formation of ETBE. For instance, A-15, A-35, and ZSM-5 zeolite were used by Tau and Davis 88 as catalysts in the ETBE production. The ZSM-5 zeolite used in either MTBE synthesis reported by Chu and Kuhl

23

or

ETBE synthesis reported by Tau and Davis 88 resulted the same selectivity (100%). It means that ZSM-5 is a prospective catalyst to be used in the formation of ETBE replacing MTBE as an additive gasoline. Other catalysts used in ETBE synthesis were mordenite zeolite K-2631 resin

90

89

and Lewatit

. The Lewatit K-2631 resin resulted the higher activation energy than A-15

reported by Ancillotti et al 91 and Dowex M-32 resin reported by Kaitale et al 92. Even, this resin resulted the higher activation energy than the one over A-15 used in MTBE synthesis reported by Gicquel and Torck 93 and Ancillotti et al 91.

2.

Oligomerization of isobutylene Isobutylene produced over many reactions as above explained can be utilized in many

processes. One of the isobutylene derivatives is isooctene obtained over dimerization, trimerization, and oligomerization of isobutylene using various catalysts such as resins, substrates, and zeolites under different conditions as presented on Table 1 and Table S1.

5

Several catalysts used in dimerization of isobutylene were Ni-Al2O3 94, Lewatit K-2631 95, USY, beta, mordenite, ZSM-5, ferrierite (FER)

96-98

, nickel complex

99

, A-15

100, 101

, indion 125 resin,

and indion 130 resin 102. Trimerization is a reaction between diisobutylene (DIB, also known as isooctene) and isobutylene to produce triisobutylene (TIB) where the carbon atoms are between C8 and C12. The catalysts used in TIB synthesis were WOx/ZrO2

103

, nickel

99

, sulfated titania

104

, and cation

exchange resins such as A-15 101, A-35, A-31, and A-DT 105-107 Other catalysts used in this synthesis were zeolites such as USY (Si/Al ratio of 30), beta (Si/Al ratio of 9 – 19), mordenite (Si/Al ratio of 12.5), ZSM-5 (Si/Al ratio of 25), and FER (Si/Al ratio of 10) catalyst

80

97, 98, 105-107

. The high selectivity to trimers can be obtained over WOx/ZrO2 as a

calcinated at 550 – 800 oC

103

with stable process

80

. The selectivity of TIB was

increased with increased Ni loading 94, lower Si/Al ratio 97, higher reaction temperature 98, lower space velocity 107, and increased isobutylene conversion 105. However, the selectivity of TIB was decreased with higher pressure 101 and longer reaction time 97. Teraisobutylene (TEB) was obtained from oligomerization reaction between TIB and isobutylene with the number of carbon atoms of ˃ C12. The catalysts utilized in oligomerization of isobutylene were Na3HSiW12O40/SiO2 108, A-15, A-35 with acid capacity of 5.2 mequiv H+/g, A-DT with acid capacity of 3.1 mequiv H+/g, A-31 with acid capacity of 4.8 mequiv H+/g 105 and Dowex 50 with acid capacity of 5.02 mmoleqv H+/g 109. Zeolites such as USY (3D), Y (3D), beta (3D), mordenite (1D), ZSM-5 (3D) and FER (2D) with surface area of 400 – 720 m2/g 109

97, 98, 105,

were successfully utilized in the oligomerization reaction.

6

7

Table 1. Dimerization, trimerization, and oligomerization of isobutylene over different catalysts 94, 96-105, 108-111. Method

Phase

Dimerization Dimerization Dimerization Trimerization Trimerization Trimerization Trimerization Oligomerization Oligomerization Oligomerization

L L L L G–L G–L -

Temp [oC] 65 – 110 60 60 – 120 70 40 – 100 30 – 110 50 – 110 20 -

Time [h] 2 16 ~ 100 1 – 12 10 2 24 90 40

X [%] 5 – 95 8 100 – 10 100 100 91 ˃ 90 41 100 100

Sel [%] 94 – 16 ˃ 99 25 – 100 75 – 80 70 66 80 ~97 76 88

Catalyst

Modification

Ref.

Ni/Al2O3 Nickel Amberlyst 15 resin WOX/ZrO2 H-FER zeolite Amberlyst 35 resin Amberlyst 15 resin Na3HSiW12O40/SiO2 Sulfated titania (TiO2) Sulfated titania (TiO2)

The incipient impregnation method Aluminum alkyl-treated nickel (1.5 mol%) Synthesized over sodium hydroxide Calcinated at 700 oC Calcination at 550 oC for 8 h to form a proton Ion exchanged over ethanol to remove water Cleaning using acetone for 70 h Modified over Na+ ion in standard method Treatment via in situ over sulfuric acid Synthesized over tert-butanol and n-butoxide

94

L = liquid; G = gas; X = Conversion; Sel = Selectivity.

8

99 100 103 98 105 101 108 111 111

Other catalysts such as MFI (silicalite) with pore diameter of 500 nm in ZMR W2O3/Al2O3

112

, H4SiW12O40/SiO2

113

modified over Na+ cation

108

, and sulfated titania

84

,

104, 111

were also used in the synthesis. Sulfated titania (TiO2) catalysts with Lowry:Brønsted ratios of 1.25 – 3.65 were modified over sulfuric acid

111

and tetra propyl ammonium bromide (TPABr)

and tetra ethyl ortho silicate (TEOS) were utilized in the silicalite membrane synthesis with the typical synthesis of TPABr : TEOS : SiO2 : H2O : KOH : NaOH = 1 : 20 : 21 : 560 – 788 : 0.9 : 3 84, 114, 115

.

The selectivity of TEB was increased with higher Ni loading 94, increased Na+ cation content 108

, and higher isobutylene concentration

107

. However, the selectivity of TEB was decreased

with lower pressure 101 and space velocity 105, 107. The selectivity obtained over dimerization is the highest selectivity to DIB among TIB and TEB with the one of more than 99% (almost 100%) using nickel as a catalyst

99

. The

dimerization of isobutylene was carried out using various catalysts such as resins, substrates, and zeolites. In Table 2 and Table S2, we summarize previously reported dimerization of isobutylene over different resins and zeolites as catalysts where polar compounds were used as additives and different solvents were applied in different processes. Some metal oxide-based catalysts in dimerization were nickel complex treated over aluminum alkyl 99, NiSO4/γ-Al2O3 116, Al2O3 modified over ammonium phosphate 117, and nickel 94, 118, 119

. Nafion SAC-13 and Nafion NR-50 synthesized over H3PW12O40.nH2O were also used

in dimerization of isobutylene

120

. The conversion was increased from 3% to 100% because of

the presence of 0.003 moles of diethyl aluminum chloride (EtAlCl2) 118. However, the selectivity was significantly decreased from 98% to less than 1%. The commercial resins utilized in

9

dimerization reaction were mostly ion exhange resins

83, 121-128

such as A-15

83, 118, 127, 129, 130

, A-

36 126, A-35 83, 110, 130, A-IR-120 118, A-IR-118 118, Dowex 50 109, and Tulsion T-63 128, 131. Selectivity enhancers such as water, methanol, TBA, MTBE, isopropanol, and 2-butanol were utilized in the synthesis to increase the selectivity of DIB 133

7, 83, 100, 110, 121, 122, 124-126, 128-130, 132,

. The high selectivity in a reactive distillation was achieved because of the synergetic effect of

distillation and reaction 123 whereas the selectivity was not affected by the different particle sizes of catalysts 124. The selectivity enhancers have important impact on the selectivity and yield of DIB. The selectivity and yield of DIB without additives were 24% and 23%, respectively 125. However, the selectivity and yield of DIB were significantly increased using methanol and tert-butyl alcohol (TBA) where the selectivities and yields were 48% and 43% for methanol and 86% and 46% for TBA, respectively,

125

. The high selectivities and yields using methanol and TBA can be

achieved because of high polarity 130 whereas these polar components significantly decreased the activity of catalyst 100, 125. The effect of polar components such as methanol, ethanol, and isopropanol was also observed by Girolamo et al

110

. The highest selectivity to DIB was obtained using isopropanol

with the one of more than 90% for 3 h under alcohol:isobutylene molar ratio of 0.2 and isobutylene content of 45 – 50%. The selectivity to DIB obtained using ethanol was higher than using methanol with the ones of more than 80% for ethanol and less than 80% for methanol

110

.

The selectivity to DIB obtained over TBA was higher than over MTBE additive with the one of 80% for MTBE and 100% for TBA at 80 oC 125.

10

Table 2. Dimerization of isobutylene to isooctene over different catalysts 94, 99, 100, 110, 116-118, 120-126, 128-131, 134.

n-pentane m-xylene 2-methyl butane Isooctane Isooctane TBA–isooctane -

Temp [oC] 60 65 – 110 60 – 120 50 – 90 60 – 90 65 – 95 44 – 75 90 -

Pressure [atm] 20 21 10 25 15 ˃ 15 -

Time [h] Up to 30 2 2 11 -

Conversion and Selectivity [%] X = 2 – 5.5; S = 100 X = ˃ 95; S = 94 X = 16 – 70; S = 93 – 95 X = 10 – 20; S = 75 – 85 X = 99.58; S = 85.71 X = ~ 60; S = 91 – 94 X = 21 – 63; S = 83 – 98 X = 90 – 99; S = 80 – 95 X = 50 – 95; S = 85 – 96 X = 44; S = 74

Isopentane

60 – 90

13 – 15

-

X = 100 – 10; S = 24 – 100

100

Isopentane Isooctane Isopentane Organic n-decane

70 – 110 30 – 50 60 – 120 60 45

15 – 18 3.5 – 4.4 14.8 6–7 6–7

1–4 Up to 12

X = 30 – 70; S = 97 X = 22 – 98; S = 20 – 86 X = 40 – 60; S = 80 – 90 X = ~ 10; S = 100 X = 100; S = 99

126

Catalyst

System

Selectivity enhancer

Solvent

Al2O3-phosphated Ni/Al2O3 Nickel complex 1 NiSO4/γ-Al2O3 Tulsion T-63 Tulsion T-63 Ion exchange resin Ion exchange resin Ion exchange resin Beta – 30 zeolite

PFR BR BR BR Batch RD BR FBR RD PFR -

Amberlyst 15

CSTR and BR

Amberlyst 36 Nafion SAC Ion exchange resin Amberlite IR-120 Nickel salt

CSTR CM CSTR FBR and BR FBR and BR

Water and TBA TBA MeOH, TBA, MTBE and 2-butanol TBA TBA -

Note : S = selectivity; X = conversion; TBA = tert-butyl alcohol; MeOH = methanol; MTBE = methyl tert-butyl ether; CBV-760 = zeolite H-sdusy powder; RD = reactive distillation; CD = catalytic distillation; PFR = plug flow reactor; CM = catalytic membrane; FBR = Fixed bed reactor; BR = batch reactor; CSTR = continuous stirred tank reactor; TR = tubular reactor.

11

Ref. 117 94 99 116 131 128 121 122 123 134

120 124 118 118

The selectivity of DIB in dimerization of isobutylene was decreased over sodium content on resin based catalysts 124, 126, 135

. However, the selectivity can be increased using the polar components

, lower temperature

concentration 126

126

128

120, 125

, larger pore size of resins

127

, and smaller isobutylene

where the good selectivity was achieved using a medium crosslinked catalyst

. Another catalyst such as zeolite is also used in dimerization of isobutylene. Zeolites used in

the synthesis were Beta 109 with Si/Al ratios in the range of 10 to 150 134, mordenite with particle sizes of 20 – 35 mesh

109

, Y (FAU topology) with Si/Al ratios of 10.8 – 48

118

, and ZSM-5

136

with Si/Al ratios of 140 – 230 118. Physical properties of catalysts such as zeolites and resins used in dimerization of isobutylene are presented on Table 3 and Table S3. Zeolite Y has the largest specific surface areas of approximately 687 – 763 m2/g with acid capacity of 0.82 mequiv H+/g and pore diameter and pore volume of 3.61 nm and 0.17 cm3/g, respectively

96, 118

. Beta and

mordenite zeolites have surface areas of 650 m2/g and 500 m2/g, with pore diameter of 8.75 nm and 3.62 nm, respectively 96, 97. ZSM-5 has a smaller surface area of about 402 – 418 m2/g with Na2O contents of 0.03 – 0.05

118

. The effect of porosity of catalysts on the conversion and

selectivity of dimerization will be discussed on the next section (Section 3.3).

12

Table 3. Physical properties of catalysts used in dimerization of isobutylene 96-98, 118, 124, 126, 128, 130, 135-138. Properties cross-linking (%) exchange capacity (eq H+/kg) porosity (cm3/g) specific surface area (m2/g) pore diameter (nm) temperature stability (oC) particle size (mm) acidity (mequiv H+/g.dry) Si/Al ratio Na2O (%) moisture content (%) apparent density (g/cm3) pore volume (cm3/g)

T-63

Y-Z

A-15

P-CT 252

15 4.9 35 40 120 0.3 – 1.2 40 -

687 – 763 3.61 0.82 10.8 – 48 0.03 – 0.1 0.1654

H 4.81 0.30 45 – 157 30 – 34 120 0.63 – 1.25 4.75 53.1 0.77 0.3

M 5.40 132 39 130 0.78 5.40 0.2 – 0.5

T = tulsion; Z = zeolite; A = amberlyst; P = purolite; Y = FAU topology; H = high; M = medium.

13

The production of isooctene and isooctane can be produced over many of the presently existing commercial technologies in the available industries as presented on Table 4. Additives such as TBA, methanol, MTBE, and water were utilized in the process as a selectivity enhancer. Ion exchange resin was widely used as a catalyst by the most of technologies. However, Dimersol technology applied by IFP company used nickel complexes as a catalyst in the process to convert linear butenes to isooctene. This catalyst was synthesized using an organo chloro aluminate ionic liquids (OCAIL). The high selectivity of 92 – 93% was obtained without additive 100, 139. The obtained isooctene can be easily separated because of poorly soluble solution. Ortho phosphoric acid used as a catalyst by Nippon Oil’s Olefin Conversion Technology resulted the higher selectivity approximately 100% without selectivity enhancer 140

. The catalyst was synthesized using PAM supported on silica

140

100,

. Another catalyst beside ion

exchange resin utilized in isooctane production was a solid phosphoric acid (SPA) used by UOP’s InAlk technology 100, 130, 141, 142. The catalyst was treated using polyphosphoric acid (PA) where water, TBA, and methanol were used in the process to increase the selectivity of isooctene or isooctane. The commercial technologies using water or pure TBA as a selectivity enhancer and ion exchange resin as a catalyst were Neste Oil and KBR’s NExOCTANE technology UOP’S InAlk technology

100, 141, 142

100, 143, 144

, and Lyondell’s Alkylate 100 technology

100

,

with

deactivation of catalysts using sodium hydroxide solution, surface areas of 33 – 53 m2/g, pore diameters of 24 – 30 nm, and acid capacities of 4.7 – 5.4 mmol H+/g 126.

14

Table 4. Commercial technologies used to produce isooctane over the hydrogenation of isobutene dimers. Commercial Technology NExOCTANE CDIsoether

Industry Neste Oil and KBR CDTECH and Snamprogetti

Selectivity enhancer Water and TBA Water/TBA and MeOH/MTBE Water, TBA, and methanol

InAlk

UOP

OCT

Nippon Oil

-

IFP Lyondell

Water and TBA

Dimersol Alkylate 100

Catalyst

Synthesis

Ion exchange resin

Isooctane properties

Ref

-

RON 99.1

MON 96.3

SG 0.70

Ion exchange resin

-

100.2

100.3

0.72

83, 100, 145

SPA or SAR

SPA treated over PA

100

100

-

100, 130, 141, 142

Ortho phosphoric acid Nickel complexes Ion exchange resin

100, 143, 144

PAM supported on silica

-

-

100, 140

Treated using OCAIL treated over NaOH

-

-

100, 139 100, 126

KBR = Kellogg Brown and Root; CDTECH = catalytic distillation technologies; OCT = Olefin Conversion Technology; SPA = solid phosphoric acid; SAR = sulfonic acid resin; PA = polyphosphoric acid; PAM = phosphoric acid monomer; OCAIL = organo chloro aluminate ionic liquids; RON = Research Octane Number; MON = Motor Octane Number; SG = specific gravity.

15

The catalytic distillation technologies (CDTECH) and Snamprogetti’s CDIsoether technology used methanol or MTBE in addition into pure TBA to increase the selectivity of product

83, 100, 145, 146

and A-15 (pore diamater of 24 nm), A-35 (pore diamater of 20 nm), A-

XN1010 (pore diamater of 5 nm), and A-XE586 (pore diamater of 25 nm) as catalysts

83

where

dimerization and etherification were partially performed. The alcohols such as methanol or MTBE have important impact on isoether technology for isooctane synthesis via hydrogenation from dimerization reaction because of high polarity, shorter removing the reaction heat, and a poor control of temperature 83. Isooctane obtained over CDIsoether technology by CDTECH and Snamprogetti had research octane number (RON) and motor octane number (MON) of 100.2 and 100.3 and specific gravity of 0.72 83. Figure 2 shows a flowscheme isooctane production from isobutylene using NExOCTANE technology adapted from KBR technology. Isobutylene, which is used in dimerization, can be derived from fluid catalytic cracking (FCC) 149, 150

40, 147, 148

, naphtha and ethane steam cracking (SC)

, oxidative dehydrogenation of n-butene (n-butene ODH)

32, 151-155

, and skeletal

isomerization of n-butenes 44, 46-48, 50-52. The FCC produced 50.5 wt.% of C4 olefins containing 15 wt.% of isobutylene C4 olefins

32

40

147

using zeolites as catalysts

using various catalysts

and n-butene ODH produced 83.3 wt.% of

151, 153, 155, 156

. Approximately, 5.76 wt.% of C4 olefins was

produced over naphtha cracker 149 and 94 wt.% of C4 olefins was produced over the low severity SC process comprising 32 wt.% of isobutylene

32, 157

. The obtained C4 olefins were affected by

different temperatures 158, 159, the types of feedstocks 159, 160, and catalysts 159, 161. Production of isooctene was processed over dimerization of isobutylene and further hydrogenated to obtain isooctane using one of the commercial technologies such as NExOCTANE technology because of a low cost technology as presented on Figure 2. Isooctane

16

used as an octane booster compound in gasoline for fuel engines has octane number of 100 and n-heptane of 0 100. Isooctane production via hydrogenation of isooctene was performed in a three-phase reactor using a mixture of 2,4,4-trimethyl 2-pentene and 2,4,4-trimethyl 1-pentene with ratio of 1:4 and total concentration of 5 or 15 mol.% in the liquid phase, cyclohexane as a solvent, and Ni/Al2O3 catalyst with surface area of 108 m2/g and pore volume of 0.37 cm3/g at 35 – 95 oC and 10 – 14 atm for 5 – 12 h

162

. The obtained activation energies were 34 kJ/mol for 2,4,4-trimethyl 1-

pentene and 49 kJ/mol for 2,4,4-trimethyl 2-pentene

162

. Approximately, 60 – 65 kJ/mol of the

activation energy was obtained from etherification for 2,4,4-trimethyl-pentenes

163

using

methanol 164. The NExOCTANE technology was carried out in a fixed bed reactor in the liquid phase using water or TBA as a selectivity enhancer

100

and ion exchange resin as a catalyst with the

feedrate of approximately 13,700 lb/h at 57 oC and 30 atm feed component

119

119, 144

. Isopentane was an inert in the

. The catalysts used in the NExOCTANE technology had an excellent

performance and longer activity than others for saving cost. The NExOCTANE technology was operated and developed by Neste Oil and KBR technology. The NExOCTANE technology is chosen as a promising technology in the future for isooctane production and suitable for many processes such as FCC refinery, C4 olefins production, and isobutylene dehydrogenation process because it has many advantages. First of all, it is a low cost effective technology and appealing in the conversion of existing MTBE facilities. The second advantage, dimerization process and hydrogenation process were separated each other so the production of isooctene can be carried out without hydrogenation process in the NExOCTANE technology. It also produces the best

17

quality commercial product, high availability, and long catalyst age. In addition, it can be utilized by producer to the use of existing hydrogenation units or to mix isooctene into the gasoline. The mechanism of isooctane production obtained over dimerization of isobutylene is followed by hydrogenation process as presented on Figure 3 where the hydrogenation units were utilized to remove sulfur content to be less than 5 wppm with a reaction heat of about – 30,000 kcal/kmol because of high exothermic

165

. Catalysts are required in hydrogenation process to

fasten the reaction rate and shorten the reaction time and the catalysts mostly used in hydrogenation process were synthesized over nickel. The poisons on isobutylene have important impact on the catalytic reaction in the hydrogenation process 165.

3. Effect of heterogeneous catalysts properties in the oligomerization of isobutylene 3.1 Effect of different catalysts such as resins and zeolites Catalytic oligomerization of isobutylene to obtain DIB, TIB, and TEB was carried out using different catalysts such as resins and zeolites. Comparison of the effectiveness between resins such as A-15 and A-35 and zeolites such as Beta zeolite and ferrierite zeolite as catalysts in trimerization reaction was reported by Yoon et al

106

. The zeolites were calcinated to the proton

form and the resins comprised sulphonic acids. The conversion of isobutylene obtained over resins was very stable with the one of 100% up to 600 h where the performance of A-35 was better than A-15 because of higher acidity. The conversion obtained using A-35 was decreased over higher weight hourly space velocity (WHSV) for 20 h

107

. However, the conversion of

isobutylene obtained over zeolites showed the worse performance compared with those over resins. The one for Beta zeolite was 100% – 90% within 300 h and 100% – 80% for ferrierite zeolite up to 150 h 106.

18

Moreover, the selectivities of TIB obtained over resins were higher than the ones obtained over zeolites. The highest selectivities of TIB were approximately 80% for A-35, 75% for A-15, 70% for ferrierite zeolite, and 60% for Beta zeolite 106. The selectivity of TIB obtained using A35 and ferrierite zeolite was decreased because of higher WHSV

107

. For instance, the

selectivities of TIB obtained over A-35 were more than 70% with WHSV of 10 h-1, 60% with WHSV of 30 h-1, and less than 50% with WHSV of 50 h-1

107

. The advantages of resins used in

oligomerization as catalysts were long lifetime and high selectivity of TIB

106

. The deactivation

rate of resins was rapidly decreased with lower temperature as reported by Hauge et al 109. The conversion and selectivity of TIB in trimerization of isobutylene obtained using mordenite zeolite were smaller than those obtained using resins such as A-35, A-15, and A-DT where A-15 produced the highest conversion and selectivity of TIB because of the highest acid capacity 105. The resins were ion exchanged over ethanol to remove water content and mordenite zeolite was calcinated at 550 oC for 8 h. The conversion and selectivity of TIB obtained using mordenite zeolite were 70% and 16%, respectively, and the conversion and selectivity of TIB achieved using resins were 71% and 50% for A-DT, 90% and 64% for A-15, and 91% and 66% for A-35, respectively 105. The comparison between resins such as A-15 and Dowex 50 and zeolites such as ZSM-5, mordenite, beta and Y in oligomerization of isobutylene was reported by Hauge et al

109

. The

resins were slowly deactivated so that the conversion was slowly decreased with the one of from 70% to 35% for A-15. However, the zeolites were rapidly deactivated so that the conversions were sharply decreased with the ones of from more than 90% to less than 10% for zeolite Y, from more than 85% to less than 10% for Beta zeolite, from more than 60% to less than 10% for ZSM-5 and mordenite zeolite

109

. The A-15 showed a good stability

109

whereas the selectivity

19

over zeolites was low because of small pore size and strong acidity 106. The advantage of zeolites as catalysts is their regenerability over simple calcination. The shorter lifetime and the lower selectivity on zeolite catalysts in oligomerization reaction can be improved with meso- and macroporous sizes 106.

3.2 Effect of Topology (1-D versus 3-D) Topology has very important impact on catalyst design for dimerization, trimerization, and oligomerization of isobutylene. Various zeolite topologies were used in oligomerization of isobutene. Yoon et al

98, 106

studied the effect of different topologies in oligomerization of

isobutylene using Beta zeolite (BEA) zeolite (FER)

106

106

with three dimensional channels (3D)

with two dimensional channels (2D)

166

166

, ferrierite

, and mordenite zeolite (MOR)

98

with

one dimensional channels (1D) 166, The BEA and FER topologies were calcinated at 550 oC for 8 h and reactivation of aged zeolites was performed over calcination at 400 oC for 10 h

106

. The

MOR topology was heated at 300 oC for 10 h to remove water content 98. The selectivity to TEB obtained over BEA (3D) was the highest among the others and the one obtained over FER (2D) was higher than the one obtained over MOR (1D). The selectivity to TEB was in the order of BEA > FER > MOR because the dimension of BEA (3D) was the largest among FER (2D) and MOR (1D) and the dimension of FER (2D) was larger than the one of MOR (1D). The highest selectivity to TEB for BEA (3D) was more than 30% and the one for FER (2D) was approximately 20% 106. However, the selectivity to TEB obtained over MOR (1D) was less than 5%

98

. The BEA (3D) was slowly deactivated and had the highest selectivity of

TEB because of the largest channel system

106

. The higher selectivity of TEB can be achieved

over larger channel system.

20

Moreover, the conversion obtained over BEA (3D) was higher than over FER (2D) where the ones were approximately 99% for BEA (3D) and 83% for FER (2D) for 150 h

106

. The

conversion obtained over MOR (1D) was approximately 20% with the fastest reaction time of 12 h

98

. The zeolites with one dimensional channel (1D) such as mordenite zeolite (MOR)

SAPO-11 zeolite (AEL)

106

98

and

were rapidly deactivated because of the smallest channel systems.

Therefore, the conversion of isobutylene obtained using these zeolites was sharply reduced. The effect of topology (1-D versus 3-D) in trimerization of isobutylene using mordenite zeolite (MOR) with one dimensional channel (1D) and ZSM-5 zeolite (MFI) with three dimensional channels (3D) was reported by Yoon et al 98. The conversion on the MOR (1D) was sharply decreased approximately from 90% to 20% for 12 h with deactivation of approximately 6.4% per hour whereas the one on MFI (3D) was slowly decreased approximately from 60% to 40% for 12 h with deactivation of approximately 1.8% per hour

98

. The selectivity of TIB

obtained over MFI (3D) was higher than over MOR (1D) with the one of approximately 15% for MFI (3D) and less than 10% for MOR (1D)

98

. The MFI (3D) was slowly deactivated and had

higher selectivity of TIB because of the larger channel systems. The recent MFI synthesis was carried out using isopropylamine as an organic structure directing agent (OSDA) and 2-propanol as co-solvent 167. The (70%) selectivity of TIB obtained over the FER (2D) was much higher than the (10%) one obtained over MOR (1D) with the steady conversion (100%) for up to 12 h because of the larger channel system on FER (2D) 14.13, and c = 7.49 Å

168

98

. The FER (2D) had unit-cell dimensions a = 19.16, b =

and was also widely utilized as a catalyst in many applications as

previously reported by some researchers

15, 169-187

. The recent synthesis of FER topology zeolite

21

was carried out using various organic structure directing agents (OSDAs) at 60 – 200 oC with the fastest reaction time of 0.5 h and the longest reaction time of 336 h 188-200. The effect of topology (1-D versus 3-D) in trimerization of isobutylene using the BEA (3D) with Si/Al ratio of 12.5 and the MOR (1D) with Si/Al ratio of 12.5 was also studied

97

. In this

reaction, both catalysts were heated at 300 oC for 10 h to remove water content. The obtained result was the same with the result obtained by Yoon et al

98

where the zeolites having larger

dimensional channel produced the higher selectivity and the stable conversion during the reaction time. The selectivity of TIB obtained over the BEA (3D) was more than 60% for 12 h with the conversion of 100% up to 25 h 97. However, the smaller selectivity of TIB obtained over the MOR (1D) was less than 10% for 12 h and the conversion was rapidly decreased approximately from 90% to 20% for 13 h 97. The catalytic performance of the Beta zeolite (BEA) with Si/Al ratio of 12.5 was compared with the USY zeolite (FAU) with Si/Al ratio of 30 97 in trimerization of isobutylene where these zeolites have the same dimensional channel (3D) and the same pore size (12 membered rings) 166

. A result showed that the BEA (3D) had the remarkable performance compared with the FAU

(3D). The selectivity to TIB obtained over the BEA (3D) was higher than the one obtained over the FAU (3D). Moreover, the conversion on the BEA (3D) was very stable for 25 h whereas the one on the FAU (3D) was slowly decreased with the reaction time of 20 h. The selectivities of TIB were more than 60% for the BEA (3D) and more than 40% for the FAU (3D). The conversions were 100% for the BEA (3D) and from 100% to about 85% for the FAU (3D)

97

.

The FAU (3D) was ion exchanged over AgNO3 and Cu(NO3)2.3H2O at 25 oC for 24 h 201. The conversion and selectivity obtained over the BEA (3D)

97, 106

both in the trimerization

reaction and in the oligomerization reaction were the highest among the ones obtained over

22

MOR (1D) 97, 98, FER (2D) 106, FAU (3D) 97, and MFI (3D) 98. Therefore, the BEA (3D) was the most potential catalyst to use in the oligomerization of isobutylene to produce TIB and TEB because of the highest selectivity, high stability, low deactivation, easy regeneration, and stable conversion

97

. The performance of the BEA (3D) was decreased with the higher Si/Al ratio

97

.

Further improvement of BEA (3D) can be carried out over microwave synthesis on a borosilicate glass substrate

202

, using choline chloride/urea mixture as a deep eutectic solvent

203

, and the

incorporation of Mn into the framework of Beta zeolite 204. Even though, the MOR (1D) produced the lower selectivity of TIB and TEB than those produced over other catalysts because of the smallest dimensional channel. The MOR (1D) had a remarkable performance in dimerization of isobutylene to produce DIB, well known as isooctene. Dimerization of isobutylene was carried out using three different catalysts such as the MOR (1D), the FAU (3D), and the BEA (3D) with the same pore size of 12 member rings

97

.

The selectivities of DIB obtained over the BEA (3D) and the FAU (3D) were less than 10% and 40%, respectively. However, the selectivity of DIB achieved over the MOR (1D) was the highest selectivity of DIB with the one of approximately 90%

97

. The MOR can be dealuminated over

nitric acid (HNO3) 205, microwave 206, 207, and chemical reagents 208. The effect of topology (2-D versus 3-D) in dimerization of isobutylene using ferrierite zeolite (FER) with two dimensional channels (2D) and ZSM-5 zeolite (MFI) with three dimensional channels (3D) was also observed by Yoon et al 98. Both catalysts were calcinated at 550 oC for 8 h then converted into the proton form and then heated at 300 oC for 10 h. Even though, the FER (2D) showed a very stable conversion with the one of 100% up to 12 h, the selectivity of DIB obtained over FER (2D) was much smaller than over MFI (3D) because of the smaller dimensional channel. The selectivities of DIB obtained over FER (2D) and MFI (3D)

23

were more than 10% and 80%, respectively, after 10 h reaction time

98

ZSM-5 zeolite was carried out using both different organic templates

209-213

. The recent synthesis of and without organic

templates 214.

3.3 Effect of Porosity A better thermal stability, the high micropore surface area with pore diameter of 0.3 – 1.5 nm

215

and strong acid sites

216

, lead to zeolites are mostly used as catalysts in many reactions.

The different pore sizes of zeolite have significant impact on conversion and selectivity in dimerization, trimerization, and oligomeerization of isobutylene. For instance, Beta zeolite (BEA) with 12-member rings (MR-12) and ferrierite zeolite (FER) with 10-member rings (MR10) were applied in oligomerization of isobutylene to produce TEB with the reaction time of 600 h 106. The selectivity of TEB obtained using BEA (MR-12) was higher than using FER (MR-10) because the pore size of BEA (MR-12) was larger than the pore size of FER (MR-10). The highest selectivities of TEB were approximately 20% for FER (MR-10) and more than 30% for BEA (MR-12) 106. Moreover, deactivation in BEA (MR-12) was slower than the one in FER (MR-10) because of the larger pore size. It leads to the concentration on the FER (MR-10) was rapidly decreased. The conversion of isobutylene were approximately decreased from 100% to 90% for BEA (MR12) and from 100% to 80% for FER (MR-10) 106. The pore size had also significant impact on dimerization of isobutylene to produce DIB. Mordenite zeolite (MOR) with 12-member rings (MR-12), ferrierite zeolite (FER) with 10member rings (MR-10), and ZSM-5 zeolite (MFI) with 10-member rings (MR-10) were compared in this reaction to observe the catalytic performance of zeolites 98. The same trend was

24

obtained as in trimerization of isobutylene where the zeolites having the larger pore size showed the better performence. The selectivity to DIB on MOR (MR-12) was the highest among the one on FER (MR-10) and the one on MFI (MR-10) because of the largest pore size. The selectivities to DIB were approximately 90% for MOR (MR-12), 83% for MFI (MR-10), and less than 20% for FER (MR-10) after the reaction time of 10 h 98.

3.4 Effect of acidity Acidity affected the isooctene production obtained from dimerization, trimerization, and oligomerization of isobutylene using various catalysts such as substrates, resins, and zeolites. At Al2O3 catalyst synthesized over 1 – 6 % of ammonium phosphate, Lewis acid sites had more important impact on dimerization of isobutylene than Brønsted acid sites because of enough acid strength with high selectivity

117

. The higher catalysts stability was achieved with the higher

Lewis/Brønsted sites ratio in oligomerization of isobutylene using sulfated titania sol (TiO2) catalyst 111. The total acidities with Lewis/Brønsted sites ratios of 3.65, 1.82, and 1.25 were 280, 105, and 160 mmol/g, respectively. The conversion of isobutylene and selectivity of C 8 – C12 olefins using sulfated TiO2 synthesized over ammonium sulfate for 4 h were 100% and 88%, respectively, with the reaction time of 40 h 111. The effect of acidity on resins was reported by Yoon et al

105

in trimerization of isobutylene

using A-35, A-15, and A-DT resins. These resins were ion exchanged using ethanol to remove water content. The acid capacities on A-35, A-15, and A-DT resins were 5.2 mequiv.H+/g, 4.7 mequiv.H+/g, and 3.1 mequiv.H+/g. The conversions and selectivities to TIB obtained using the resins were 91% and 66% for A-35, 90% and 64% for A-15, and 71% and 50% for A-DT,

25

respectively

105

. The A-35 resin produced the highest conversion and selectivity because it has

the highest acid capacity 105. Zeolite was one of the most important catalysts in the industry because of its tunable acid sites 215, 216. The acidity of zeolites modified by nickel salts such as NiCl2, NiSO4, and NiCO3 has significant impact on catalytic activity, stability, and selectivity of zeolites 96. For instance, more active and selective to dimerization reaction was achieved over Y zeolite modified over NiCO 3 96

. The activity was significantly reduced over converting the most of Brønsted acid sites to

Lewis acid sites

217

and the unsaturated compounds were formed by the strong Lewis acid sites

(with the highest number of 0.923 mmol/g at 20 oC)

96

. The high selectivity to DIB in

dimerization reaction was probably enhanced because of Lewis acid sites having enough acid strength. However, the Brønsted acid sites were not observed

117

. Therefore, both Brønsted acid

sites and Lewis acid sites have important impact on dimerization reaction to isooctene. The effect of acidity on membrane zeolite beta (MZB) with SiO2/Al2O3 ratios of 60, 90, and 120 was also observed on the resulted selectivity to DIB in oligomerization of isobutylene at 100 – 150 oC

218

. The MZB-60 had the highest acid capacity among the MZB-90 and the MZB-120.

The acid capacities on MZB-60, MZB-90, and MZB-120 were 4.6 x 10-3, 3.07 x 10-3, 1.82 x 103

, respectively. The obtained selectivities to DIB were 92% for MZB-60, 84% for MZB-90, and

74% for MZB-120

218

. The higher acidity produced the higher selectivity. The MZB-60 resulted

the highest selectivity because of the highest acid capacity. The best conversion of isobutylene and high selectivity to TIB in trimerization of isobutylene using Beta-25 zeolite (SiO2/Al2O3 = 25) were achieved because of high Lewis:Brønsted site ratio with the selectivity to TIB of more than 50% at isobutylene WHSV of

26

10 h-1 for up to 100 h

97

. The Beta-25 zeolite had the better catalytic activity and higher

conversion of isobutylene than Beta-18 zeolite (SiO2/Al2O3 = 18) and Beta-38 zeolite (SiO2/Al2O3 = 38) with the high concentration of Lewis acid site

97

.

The acidity on commercial zeolites with different topologies such as USY zeolite (FAU) and Beta zeolite (BEA) influenced the obtained conversion of isobutylene. The numbers of Brønsted acid site and Lewis acid site on Beta zeolite were 247 µmol/g and 104 µmol/g at 200 oC, 204 µmol/g and 80 µmol/g at 300 oC, and 144 µmol/g and 67 µmol/g at 400 oC, respectively, with deactivation constant of 0.2060/h 219. However, the numbers of Brønsted acid site and Lewis acid site on USY zeolite were lower than on beta zeolite. The ones were 98 µmol/g and 61 µmol/g at 200 oC, 14 µmol/g and 29 µmol/g at 300 oC, and 0 µmol/g and 27 µmol/g at 400 oC, respectively, with deactivation constant of 0.6752/h. The conversion of isobutylene on USY zeolite was dramatically decreased from 43% to 5% for 30 h and the one on Beta zeolite was decreased from more than 45% to 10% for 30 h

219

. The deactivation of Beta zeolite was slower than USY

zeolite because of higher strength of acidity and lower deactivation constant

219

. The fast

deactivation on zeolites can be caused by blocking of the inner pores because of large oligomerization 109, 218.

3.5 Isooctane production via hydrogenation Another solution as replacing MTBE is isooctane derived from dimerization of isobutylene after hydrogenation using resin or zeolite as a catalyst in a MTBE reactor

132, 162, 165, 220, 221

. The

conditions of process such as temperature and pressure in the MTBE synthesis can be applied to isooctane process

146

. Therefore, the existing MTBE technology can be utilized for the

production of isooctane without significant changes 162. Dehydrogenation technology of paraffins

27

is an innovative technology to produce isooctane 146. Isooctane is considered as an octane booster compound because of high motor octane number (MON) and research octane number (RON) of 100 146, 162. However, the MON and RON for isooctene were 89 and 100, respectively 130. The hydrogenation kinetics of isooctene (2,4,4-trimethyl-1-pentene and 2,4,4-trimethyl-2pentene) carried out to obtain isooctane (2,2,4-trimethylpentane) in a three-phase reactor at 35 – 95 oC and 9.8 – 39 atm was reported by Lylykangas et al

162

. In this synthesis, hydrogenation

study was performed in a three-phase Robinson-Mahoney reactor in the liquid phase using cyclohexane as a solvent and Ni/Al2O3 as a catalyst where the nickel-alumina catalyst had a high activity in the hydrogenation of isooctene process. Pd/Al2O3 can be also used as a catalyst in hydrogenation process

119, 222

. The molar fraction of 2,4,4-trimethyl-1-pentene was 0.673

162

.A

heat of the reaction in hydrogenation process was exothermic with the one approximately – 109.55 kJ/mol using RSTOIC Aspen model 119. Development of isooctane production over catalytic distillation process via hydrogenation process of isooctene was carried out by Goortani et al

119

. The reduced capital and operating

costs, beneficial production of higher efficiency, and decreased waste and recycle streams were achieved because of a combination of two or more unit processes to be single unit process. The catalytic distillation process was a green technology that was utilized to combine the functions of chemical reactor (catalytic reaction) and distillation column (separation of product from reactant) into a single unit process 222, 223 and produced higher conversion and selectivity 119. Composition of a mixture of isooctene to produce isooctane was adapted from the previous study as reported by Karinen et al

224

in equilibrium isomerization reaction of isooctene

production at 50 – 110 oC using methanol 225 and A-35 224 and Smopex-101 225 as catalysts where isooctene production was affected by isomerization of linear butenes using tert-butyl alcohol

28

and ZSM-5 zeolite synthesized over nickel as a catalyst

226

.

The formation of isooctene can be also carried out in a liquid phase using sulfonic acid resin

130

.

(TBA) as a selectivity enhancer

133

The data of isooctene production can be further applied for the development of isooctane production 162.

4.

Perspective on potential zeolite topologies for dimerization Dimerization of isobutylene to isooctene over various zeolites as catalysts was previously

reported by several researchers as presented on Table 5 and Table S4. Brown et al

227

studied

dimerization of isobutylene using wavelet shrinkage denoising combinated with time-varying flexible least squares over ZSM-5 zeolite with Si/Al in the range of 16 to 17 in a reactor at 100 o

C for 2 h. The activation energy of dimerization of isobutylene was 48.3 kJ/mol – 267 kJ/mol

with the average activation energy of 112 kJ/mol 227. Koskinen et al 228 reported isooctene synthesis using CO2 and propane as solvents in reaction medium over ZSM-5 and ZSM-23 zeolites as catalysts at 100 oC and 49 – 89 atm for 120 – 200 h. The ZSM-5 catalyst with Si/Al ratio of 23.8 had the total acidity of 380 µmol/g and the ZSM23 catalyst with Si/Al ratio of 19.7 had the total acidity of 690 µmol/g. The initial conversion obtained over ZSM-23 catalyst was more than 80%. However, the conversion obtained using CO2 as a solvent was approximately 56% for up to 200 h. It was higher than the one obtained using propane as a solvent with the conversion of 32% for up to 120 h. Stability of ZSM-23 was longer over CO2 solvent than over propane solvent

228

. Further improvement of ZSM-23 (MTT)

can be performed by microwave synthesis and alkaline treatment

229-232

stability of ZSM-23 (MTT) in hot water was affected by La and Ce

233

where hydrothermal

. The highest conversion

29

of isobutylene obtained using propane solvent and ZSM-5 catalyst was approximately 72% at 49 atm with WHSV of 25 h-1 for 1 min 228.

30

Table 5. Dimerization over zeolite based catalysts 96-98, 109, 218, 219, 227, 228, 234-236. Zeolite MZβ-60 MZβ-90 Beta USY USY HY-700 Mordenite ZSM-5 Mordenite USY MZβ-120 MSU-S/WBEA

Topology

MR

D

Si/Al

Phase

BEA BEA BEA FAU FAU FAU MOR MFI MOR FAU BEA BEA

12 12 12 12 12 12 12 10 12 12 12 12

3 3 3 3 3 3 1 3 1 3 3 3

30 45 45.7 2.6 30 1.4 12.5 25 12.5 30 60 43.2

Gas Gas Gas Gas Liquid Liquid Liquid Liquid Liquid Liquid Gas Gas

Surface area [m2/g] 455 396 811 867 750 382 500 425 500 720 237 718

Solvent n-butane n-butane n-butane n-butane n-butane n-butane -

Temp [oC] 100 – 150 100 – 150 60 60 70 70 40 – 100 40 – 100 40 – 100 40 – 100 100 – 150 60

Time [h] 96 96 30 30 20 6 Up to 25 1 – 12 1 – 12 Up to 25 96 30

Results [%] X = 66; S = 92 X = ˃ 30; S = 84 X = 47 – 10; S = 86 – 99 X = 43 – 5; S = 100 X = 85.3; S = 59.6 X = 74.1; S = 50.6 X = 90 – 20; S = ~ 90 X = 60 – 40; S = ~ 80 X = 90 – 20; S = ~ 90 X = 100 – 90; S = > 30 X = ˃ 30; S = 74 X = 55 – 35; S = 70 – 80

MR = member ring; D = dimension; X = conversion; S = selectivity; MZ = membrane zeolite; MSU-S/WBEA = meso structured aluminosilicates of beta seed.

31

Ref. 218 218 219 219 234 235 97 98 98 97 218 219

A novel reactor design namely catalytic membrane reactor utilized in oligomerization reaction in a gas phase over different Beta zeolites as catalysts was proposed by Torres et al

218

.

The Beta zeolites used in this method were membrane zeolite Beta-60 (MZB-60) with Si/Al ratio of 30, membrane zeolite Beta-90 (MZB-90) with Si/Al ratio of 45, and membrane zeolite Beta120 (MZB-120) with Si/Al ratio of 60. Aerosil 300 as a silica source, potassium hydroxide, and tetra ethyl ammonium hydroxide (TEAOH) as an OSDA were used in the synthesis. The typical synthesis had the molar ratio of Na2O : K2O : Al2O3 : SiO2 : (TEA)2O : H2O = 3.2 : 1.62 : 1 : x : 22.5 : 1080 with value of x = 60 for MZB-60, x = 90 for MZB-90, and x = 120 for MZB-120. The zeolites were ion exchanged by NH4NO3 solution. The catalytic reaction was carried out at 100 – 150 oC for 96 h. The conversion of isobutylene and selectivity of isooctene in the reaction using MZB-60 at 110 oC were 66% and 92%, respectively, and the selectivities on MZB-90 and MZB-120 were 84% and 74%, respectively with WHSV of 53 h-1 218. The catalytic performance of commercial zeolites such as Beta zeolite (Si/Al ratio of 45) and USY zeolite (Si/Al ratio of 2.5) were compered with mesoporous aluminosilicate of Beta zeolite (MSU-S/WBEA) and non-mesoporous aluminosilicate of Beta zeolite (MSU-S/SBEA) in oligomerization of isobutylene at 60 oC and 1 atm for 30 h 219. Beta seeds (Si/Al ratio of 44) was synthesized using tetra ethyl ammonium hydroxide (TEAOH) solution as an OSDA (organic structure directing agent) and dried at 100 oC for 6 h. Tallow tetra amine and HCl were used to obtain MSU-S/WBEA and digestion of beta seeds was carried out in the absence of surfactant as a mesoporous template at 150 oC for 20 h to obtain MSU-S/SBEA. The initial conversions were approximately 55% on MSU-S/WBEA, 51% on MSU-S/SBEA, 47% on Beta zeolite, and 43% on USY zeolite at 60 oC with isobutane/isobutene mass ratio of 50 wt./50 wt. At the same condition, the highest selectivities of isooctene on MSU-S/WBEA, MSU-S/SBEA, Beta, and USY zeolite were

32

approximately 80%, 79%, 99%, and 100%, respectively

219

. The selectivity of isooctene on

MSU-S/SBEA was stable because of the strong Brønsted acid sites. The commercial USY zeolite (Si/Al ratio of 30) was also treated over AlCl 3 and heated at 300 oC for 10 h

234

. The synthesis was carried out using n-butane solvent with isobutylene/n-

butane mass ratio of 50/50 (wt.%/wt.%) at 70 oC and 14.8 atm for 20 h. The USY zeolite and AlUSY zeolite (USY treated by AlCl3) were used in the synthesis as catalysts. The conversion and selectivity on USY zeolite were 85.3% and 59.6 wt.%, respectively, and the conversion and selectivity on Al-USY zeolite were 99.2% and 21 wt.%, respectively. The conversion on AlCl3 treated zeolite was higher than on commercial zeolite because of higher Lewis acid site/Brønsted acid site ratio. The commercial zeolite had Lewis acid site/Brønsted acid site ratio of 0.31 with 0.34 for Al-USY zeolite 234. USY, Beta, and mordenite zeolites were also used in dimerization reaction by Yoon et al 97. The synthesis was carried out in a liquid phase using n-butane as a solvent at 40 – 100 oC and 14.8 atm. Beta zeolite utilized was Beta-25 zeolite (SiO2/Al2O3 = 25) with surface area of 680 m2/g 97. The conversion on mordenite zeolite was reduced from approximately 90% for less than 5 h to approximately 20% for more than 10 h with the highest selectivity to DIB of approximately 90 wt.% for 12 h. Approximately, 100% of the highest conversion on USY zeolite was obtained for less than 5 h and the selectivity was more than 30 wt.% for 12 h. However, the conversion on Beta-25 zeolite was constant with the conversion of 100% up to 25 h. The selectivity to DIB was approximately 10 wt.% for 12 h 97. Yaocihuatl et al 96 utilized Beta (with Si/Al = 75), Y (with Si/Al = 2.6), and mordenite (with Si/Al = 45) zeolite for isooctene synthesis in a gas phase to reduce the effect of diffusion. These zeolites were modified over nickel salts such as NiCl2, NiSO4, and NiCO3 with different

33

concentrations (0 – 16.5 wt.%) and heated at 100 oC for 12 h and then calcinated at 500 oC for 5 h. Surface area of the zeolites were reduced after treated by the nickel salts. For instance, surface areas of beta, Y, and mordenite were 650, 660, and 500 m2/g, respectively. However, the ones were reduced to 437.5, 404.4, and 394.1 m2/g, respectively, after treated by NiSO4.6H2O. The highest conversion and selectivity to DIB obtained using Y zeolite (particle size of 0.3 mm) with 9% of nickel load at 25 oC were 0.7% and 0.98%, respectively 96. The more active and selective to dimerization process was obtained using Y zeolite synthesized over NiCO3. Dealumination of HY zeolite was carried out over calcination at 500 oC (HY-500), 600 oC (HY-600), and 700 oC (HY-700) for 12 h and then heated at 300 oC for 10 h

235

. The catalytic

reaction was performed in a liquid phase using n-butane solvent with isobutylene/n-butane mass ratio of 50/50 (wt.%/wt.%) at 70 oC for 6 h. The resulted conversions were 83.1% for HY zeolite, 91% for HY-500 zeolite, 98.4% for HY-600, and 74.1% for HY-700 and the obtained selectivities of isooctene for HY, HY-500, HY-600, and HY-700 zeolites were 32.5%, 26.3%, 16%, and 50.6%, respectively, with WHSV of 10 h-1 at 70 oC and 14.8 atm for 6 h 235. One of the potential catalysts in trimerization of isobutylene to isooctene is ferrierite zeolite (FER) because of high conversion and selectivity with the selectivity to DIB of more than 70 wt.% and the stable conversion of 100% for 12 h

98

. Because of the remarkable performance,

ferrierite zeolite was also used in dimerization of isobutylene. Comparison between resins and zeolites as catalysts used in oligomerization of isobutylene in a liquid phase using n-butane solvent at 30 – 70 oC and 10 atm was reported by Hauge et al 109. The resins used were A-15 and Dowex 50 and the zeolites used were H-Y, Beta, ZSM-5, and mordenite. The resins were heated in situ at 107 oC for more than 1 h and the zeolites were calcinated using dry air at 500 oC and 1 atm for 8 h. The conversions of the all catalysts were

34

decreased corresponding to the longer synthesis time. Even though, the A-15 resin had the best stability compared zeolites. A shortage of A-15 resin was smaller initial activity and lower initial conversion than Beta and H-Y zeolites. The initial conversions on Beta and H-Y zeolite were more than 80% and 90%, respectively. However, the one on A-15 resin was approximately 70% with WHSV of 60 h-1 at 40 oC and 10 atm for up to 5 h 109. Yoon et al

98

produced isooctene using three different catalysts such as ferrierite with Si/Al

ratio of 10, ZSM-5 with Si/Al ratio of 25, and mordenite with Si/Al ratio of 12.5. The synthesis was carried out in a liquid phase using n-butane solvent with WHSV of 2.5 h-1 – 20 h-1 at 40 – 100 oC and 14.8 atm. The ferrierite and ZSM-5 zeolites were calcinated at 550 oC for 8 h to create the proton form. The conversion on ferrierite zeolite was more much stable and higher than over ZSM-5 and mordenite zeolite, with approximately 20 – 90% over mordenite zeolite, 40 – 60% on ZSM-5 zeolite and 100% on ferrierite zeolite for 1 – 12 h. Nevertheless, the highest selectivity of dimer was achieved using mordenite and the smallest one was obtained using ferrierite zeolite. The selectivities on mordenite, ZSM-5, and ferrierite zeolite for 10 h were approximately 90 wt.%, 80 wt.%, and 10 wt.%, respectively. According to Table 5, the highest conversion and selectivity obtained over different zeolites in dimerization reaction of isobutylene were more than 60% 100% 90%

97

and 99%

97, 98

219

and 90%

for Beta, 100%

97, 98, 236

97

and 100%

219

for mordenite, and 100%

109

and above 80%

for USY, 98.4% 98

and 10%

98

235

98

for ZSM-5,

and 50.6%

235

for Y,

for ferrierite, respectively.

The best zeolite in dimerization process of isobutylene among the other zeolites was USY zeolite. In oligomerization of isobutylene to obtain isooctene (DIB), the USY zeolite was usually used without further purification

97, 234

and calcinated at 600 oC for 4 h

219

. Yoon et al

97

, for

35

instance, utilized the USY zeolite in trimerization of isobutylene. This zeolite was used without further purification and dehydrated at 300 oC for 10 h. However, the resulted selectivity to DIB was low with the one of less than 40 wt.% for 12 h

97

. The USY zeolite was then treated over

AlCl3 (Al-USY) as reported by Yoon et al 234 so that the conversion of isobutylene was increased from 85.3% to 99.2%. Nevertheless, the selectivity to DIB was dramatically decreased from 59.6 wt.% to 21 wt.% for 20 h

234

. The highest selectivity (100 wt.%) was achieved by Park et al

219

where the USY zeolite was calcinated at 600 oC for 4 h. However, the obtained conversion of isobutylene was sharply decreased from 43% to 5% and the reaction time was longer with time on stream of 30 h 219. Therefore, several exact strategies to optimize the performance of the USY zeolite in the reaction and synthesis of catalyst are required as presented on Figure 4. Organic template, shorter synthesis time, and low Lewis/Brønsted acid sites ratio are targeted to improve the performance of the USY zeolite in isooctane production so that dimerization of isobutylene using the USY zeolite can be further studied. The recent synthesis of USY zeolite was reported by some researchers using different organic templates 237-239. Imidazolium-based ionic liquid 239 and 3-(trimethoxysilyl)propyl hexadecyl dimethyl ammonium chloride (TPHAC)

237

can be, for

instance, utilized as a novel organic template to improve the performance of USY zeolite in dimerization of isobutylene. To obtain the higher selectivity to DIB, the low Lewis/Brønsted acid sites ratio should be applied as done by Yoon et al

234

. The shorter synthesis time leads to the

dimerization process to be more effective and promising reaction. These strategies proposed a novel development in the dimerization of isobutylene over USY zeolite by further research.

36

5.

Conclusions The production of isooctane can be obtained over many reactions such as dimerization,

trimerization, and oligomerization reactions via hydrogenation units using polar compounds such as water, methanol, TBA, MTBE, and 2-butanol as a selectivity enhancer. Dimerization is a strategic reaction to convert isobutylene, which was used as a feedstock in MTBE plant, to octane booster compound (isooctane). Heterogeneous resins, metal oxides, and zeolites were widely utilized as a catalyst in dimerization of isobutylene to isooctane using a simple and low cost effective technology. These catalysts have important impact on conversion of isobutylene and selectivity to isooctene in dimerization of isobutylene to isooctane. The conversion of isobutylene and selectivity to isooctene in dimerization reaction are also affected by several parameters such as dimensional channels, pore size (porosity on catalysts), acidity (Lewis and Brønsted acid sites), and reaction conditions. For instance, the conversion of isobutylene and selectivity to isooctene can be increased over larger dimensional channels, higher porosity, higher acid strength, lower temperatures, increased methanol:isobutylene molar ratio, and polar components. Nevertheless, the selectivity to isooctene in the dimerization reaction will be decreased over sodium content on resin based catalysts. Catalysts mostly used in the dimerization of isobutylene are commercial catalysts such as amberlyst resins, zeolite Beta (BEA), Y (FAU), ZSM-5 (MFI), mordenite (MOR), ferrierite (FER), and USY (FAU) zeolite. These zeolites are widely utilized in the synthesis as a catalyst because of easy regeneration, cheap catalysts, and more environmental friendly. The USY zeolite is one of the potential catalysts in isooctane production over dimerization reaction because of highest selectivity, better stability, high conversion, and easily regenerated catalyst. Therefore,

37

further development of the USY zeolite synthesis is required to improve the performance of the USY zeolite activity in dimerization of isobutylene to produce isooctane via hydrogenation process. Several strategies to improve the USY zeolite for commercial application in dimerization of isobutylene are expected to be developed in coming years, including organic template, shorter synthesis time, and low Lewis/Brønsted acid sites ratio.

Acknowledgement The authors would like to acknowledge the funding provided by Saudi Aramco for supporting this work through project contract number 6600011900 for the Center of Excellence in Nanotechnology at King Fahd University of Petroleum and Minerals.

38

References 1. Butler, A.; Nicolaides, C., Catalytic skeletal isomerization of linear butenes to isobutene. Catalysis today 1993, 18, (4), 443-471. 2. Sundmacher, K.; Hoffmann, U., Activity evaluation of a catalytic distillation packing for MTBE production. Chemical engineering & technology 1993, 16, (4), 279-289. 3. Li, X.; Iglesia, E., Catalytic dehydroisomerization of n-alkanes to isoalkenes. Journal of Catalysis 2008, 255, (1), 134-137. 4. Chase, J. D., Synthesis of high octane ethers from methanol and iso-olefins. In Catalytic conversions of synthesis gas and alcohols to chemicals, Springer: 1984; pp 307-322. 5. Adams, J.; Clement, D.; Graham, S., Synthesis of methyl-t-butyl ether from methanol and isobutene using a clay catalyst. Clays Clay Miner 1982, 30, (2), 129-134. 6. Morse, P., Producers brace for MTBE phaseout. Chemical & Engineering News 1999, 77, (15), 26-27. 7. Vila, M.; Cunill, F.; Izquierdo, J.-F.; Gonza´lez, J. n.; Herna´ndez, A. s., The role of byproducts formation in methyltert-butyl ether synthesis catalyzed by a macroporous acidic resin. Applied Catalysis A: General 1994, 117, (2), L99-L108. 8. Hoffman, H., Components for unleaded gasoline. Hydrocarbon Processing 1980, 60, (2), 57-59. 9. Voloch, M.; Ladisch, M.; Tsao, G., Methyl t-butyl ether (MTBE) process catalyst parameters. Reactive Polymers, Ion Exchangers, Sorbents 1986, 4, (2), 91-98. 10. Ancillotti, F.; Fattore, V., Oxygenate fuels: Market expansion and catalytic aspect of synthesis. Fuel Processing Technology 1998, 57, (3), 163-194. 11. Hatchings, G.; Nicolaides, C.; Scurrell, M., Developments in the production of methyl tert-butyl ether. Catalysis today 1992, 15, (1), 23-49. 12. Mills, G.; Ecklund, E., ALTERNATIVE FUELS-PROGRESS AND PROSPECTS. 2. Chemtech 1989, 19, (10), 626-631. 13. de Jong, K. P.; Bosch, W.; Morgan, T. D. B., Developments in gasoline reformulation and the enhancement of refinery mtbe production. In Studies in Surface Science and Catalysis, Frennet, A.; Bastin, J. M., Eds. Elsevier: 1995; Vol. Volume 96, pp 15-32. 14. Kolah, A. K.; Zhiwen, Q.; Mahajani, S. M., Dimerized isobutene: an alternative to MTBE. Chemical innovation 2001, 31, (3), 15-21. 15. Research, G. Methyl Tertiary Butyl Ether (MTBE) Global Markets to 2020 – China Dominates Global MTBE Demand While Developed Regions Look to ETBE and Ethanol as Replacement Options. http://www.gbiresearch.com/report-store/marketreports/archive/methyl-tertiary-butyl-ether-%28mtbe%29-global-markets-to-2020-chinadominates-global-mtbe-demand-while-developed-regions-look-to- (September 18, 2016),

39

16. Rehfinger, A.; Hoffmann, U., Kinetics of methyl tertiary butyl ether liquid phase synthesis catalyzed by ion exchange resin—I. Intrinsic rate expression in liquid phase activities. Chemical Engineering Science 1990, 45, (6), 1605-1617. 17. Rehfinger, A.; Hoffmann, U., Formation of Di‐isobutene, main by‐product of methyl tertiary butyl ethyl ether synthesis catalyzed by ion exchange resin. Chemical engineering & technology 1990, 13, (1), 150-156. 18. Hodge, P.; Sherrington, D. C., Polymer-supported reactions in organic synthesis. J. Wiley: 1980. 19. Adams, J.; Martin, K.; McCabe, R.; Murray, S., Methyl t-butyl ether (MTBE) production: a comparison of montmorillonite-derived catalysts with an ion-exchange resin. Clays Clay Miner 1986, 34, 597-603. 20. Baronetti, G.; Briand, L.; Sedran, U.; Thomas, H., Heteropolyacid-based catalysis. Dawson acid for MTBE synthesis in gas phase. Applied Catalysis A: General 1998, 172, (2), 265-272. 21. Quiroga, M. E.; Fígoli, N. S.; Sedran, U. A., Alternative catalysts for MTBE production. Reaction Kinetics and Catalysis Letters 1998, 63, (1), 75-80. 22. Kogelbauer, A.; Nikolopoulos, A.; Goodwin, J.; Marcelin, G., Reactant adsorption and its impact upon MTBE synthesis on zeolites. Journal of Catalysis 1995, 152, (1), 122-129. 23. Chu, P.; Kuhl, G. H., Preparation of methyl tert-butyl ether (MTBE) over zeolite catalysts. Industrial & engineering chemistry research 1987, 26, (2), 365-369. 24. Aboul-Fotouh, S. M., Production of antiknock additive in gasoline (Methyl Tert-Butyl Ether, MTBE) using zeolite catalysts. ACTA CHIMICA SLOVENICA. 2004, 51, (2), 293-304. 25. Salomón, M. A.; Coronas, J.; Menéndez, M.; Santamarı́a, J., Synthesis of MTBE in zeolite membrane reactors. Applied Catalysis A: General 2000, 200, (1–2), 201-210. 26. Hunger, M.; Horvath, T.; Weitkamp, J., Methyl tertiary-butyl ether synthesis on zeolite HBeta investigated by in situ MAS NMR spectroscopy under continuous-flow conditions. Microporous and mesoporous materials 1998, 22, (1), 357-367. 27. Mildner, T.; Ernst, H.; Freude, D.; Hölderich, W. F., Temperature-jump relaxation magic-angle spinning NMR study of the methyl tert-butyl ether reaction in a boron pentasil zeolite. Journal of the American Chemical Society 1997, 119, (18), 4258-4262. 28. Horvath, T.; Seiler, M.; Hunger, M., A comparative study of methyl-tert-butyl ether synthesis on zeolites HY, HBeta, HBeta/F and HZSM-5 by in situ MAS NMR spectroscopy under flow conditions and on-line gas chromatography. Applied Catalysis A: General 2000, 193, (1), 227-236. 29.

Trubac, R. E., Oxygenate removal in MTBE production. In Google Patents: 1989.

30. van Leeuwen, B. N.; van der Wulp, A. M.; Duijnstee, I.; van Maris, A. J.; Straathof, A. J., Fermentative production of isobutene. Applied microbiology and biotechnology 2012, 93, (4), 1377-1387.

40

31. OECD ISOBUTYLENE. (April 15, 2016),

http://www.inchem.org/documents/sids/sids/115117.pdf

32. Bender, M., An Overview of Industrial Processes for the Production of Olefins–C4 Hydrocarbons. ChemBioEng Reviews 2014, 1, (4), 136-147. 33. Pryor, P. S., Approaches to alkylation: a world review. Petroleum technology quarterly 2004, 9, 69-78. 34. Nagai, K.; Ui, T., Trends and Future of Monomer-MMA Technologies. Sumitomo Chemicals (ISSN 0387-1312) 2004, 2, 4-13. 35. Al-Megren, H.; Barbieri, G.; Mirabelli, I.; Brunetti, A.; Drioli, E.; Al-Kinany, M., Direct Conversion of n-Butane to Isobutene in a Membrane Reactor: Thermodynamic Analysis. Industrial & Engineering Chemistry Research 2013, 52, (31), 10380-10386. 36. Lee, J. W.; Kim, H. U.; Choi, S.; Yi, J.; Lee, S. Y., Microbial production of building block chemicals and polymers. Current opinion in biotechnology 2011, 22, (6), 758-767. 37. Romanow-Garcia, S.; Hoffman, H., Petroleum and its products. In Kent and Riegel’s handbook of industrial chemistry and biotechnology, Springer: 2007; pp 801-842. 38. Zwahlen, A. G.; Agnew, J. B., Isobutane dehydrogenation kinetics determination in a modified Berty gradientless reactor. Industrial & engineering chemistry research 1992, 31, (9), 2088-2093. 39. Farsi, M.; Jahanmiri, A.; Rahimpour, M., Optimal operating condition of membrane reactors to enhance isobutene production, selectivity and hydrogen production. Journal of Industrial and Engineering Chemistry 2012, 18, (5), 1676-1682. 40. Obenaus, F.; Droste, W.; Neumeister, J., Butenes. In Ullmann's Encyclopedia of Industrial Chemistry, Wiley-VCH Verlag GmbH & Co. KGaA: 2000. 41.

EIA U.S. Energy Information Administration. http://www.eia.gov/

42. Buonomo, F.; Sanfilippo, D.; Trifiro, F., Dehydrogenation reactions. Handbook of heterogeneous catalysis 1997, 5, 2140-2151. 43. Craig, R.; Spence, D., Catalytic dehydrogenation of liquefied petroleum gas by the Houdry CATOFIN and Catadiene processes. Handbook of petroleum refining processes. In New York: McGraw-Hill: 1986. 44. Houžvička, J.; Diefenbach, O.; Ponec, V., The Role of Bimolecular Mechanism in the Skeletal Isomerisation ofn-Butene to Isobutene. Journal of Catalysis 1996, 164, (2), 288-300. 45. Wattanakit, C.; Nokbin, S.; Boekfa, B.; Pantu, P.; Limtrakul, J., Skeletal isomerization of 1-butene over ferrierite zeolite: A quantum chemical analysis of structures and reaction mechanisms. The Journal of Physical Chemistry C 2012, 116, (9), 5654-5663. 46. Benitez, V. M.; Querini, C. A.; Fı́goli, N. S.; Comelli, R. A., Skeletal isomerization of 1butene on WOx/γ-Al 2 O 3. Applied Catalysis A: General 1999, 178, (2), 205-218.

41

47. Finelli, Z. R.; Fígoli, N. S.; Comelli, R. A., Isobutene production from skeletal isomerization of 1-butene on WOx/ferrierite. Catalysis letters 1998, 51, (3-4), 223-228. 48. Asensi, M.; Corma, A.; Martı, A., Skeletal isomerization of 1-butene on MCM-22 zeolite catalyst. Journal of Catalysis 1996, 158, (2), 561-569. 49. Jung, H. J.; Park, S. S.; Shin, C.-H.; Park, Y.-K.; Hong, S. B., Comparative catalytic studies on the conversion of 1-butene and n-butane to isobutene over MCM-22 and ITQ-2 zeolites. Journal of Catalysis 2007, 245, (1), 65-74. 50. Simon, M. W.; Suib, S. L.; Oyoung, C.-L., Synthesis and characterization of ZSM-22 zeolites and their catalytic behavior in 1-butene isomerization reactions. Journal of Catalysis 1994, 147, (2), 484-493. 51. Houžvička, J.; Hansildaar, S.; Ponec, V., The shape selectivity in the skeletal isomerisation ofn-butene to isobutene. Journal of Catalysis 1997, 167, (1), 273-278. 52. Seo, G.; Kim, M. W.; Kim, J. H.; Ahn, B. J.; Hong, S. B.; Uh, Y. S., Skeletal isomerization of 1‐butene on synthetic clinoptilolite zeolite. Catalysis letters 1998, 55, (2), 105-112. 53. Woo, H. C.; Lee, K. H.; Lee, J. S., Catalytic skeletal isomerization of n-butenes to isobutene over natural clinoptilolite zeolite. Applied Catalysis A: General 1996, 134, (1), 147158. 54. Lee, H. C.; Woo, H. C.; Ryoo, R.; Lee, K. H.; Lee, J. S., Skeletal isomerization of n-butenes to isobutene over acid-treated natural clinoptilolite zeolites. Applied Catalysis A: General 2000, 196, (1), 135-142. 55. Mahdi, H. I.; Irawan, E.; Nuryoto, N.; Jayanudin, J.; Sulistyo, H.; Sediawan, W. B.; Muraza, O., Glycerol Carbonate Production from Biodiesel Waste Over Modified Natural Clinoptilolite. Waste and Biomass Valorization, 1-8. 56. Pirngruber, G.; Seshan, K.; Lercher, J., Dehydroisomerization of n-Butane over Pt– ZSM5: II. Kinetic and Thermodynamic Aspects. Journal of Catalysis 2000, 190, (2), 338-351. 57. Hamid, S. B. D.-A.; Lambert, D.; Derouane, E. G., Dehydroisomerisation of n-butane over (Pt, Cu)/H-TON catalysts. Catalysis today 2000, 63, (2), 237-247. 58. Scirè, S.; Burgio, G.; Crisafulli, C.; Minicò, S., Dehydroisomerization of n-butane over HY zeolite supported Pt and Pt, Sn catalysts. Applied Catalysis A: General 2004, 274, (1), 151157. 59. Pirngruber, G.; Seshan, K.; Lercher, J., Direct Conversion of n-Butane to Isobutene over Pt–MCM22. Journal of Catalysis 2000, 190, (2), 396-405. 60.

Cheng, W.-H., Methanol production and use. CRC Press: 1994.

61. Galindo Cifre, P.; Badr, O., Renewable hydrogen utilisation for the production of methanol. Energy Conversion and Management 2007, 48, (2), 519-527. 62. Ott, J.; Gronemann, V.; Pontzen, F.; Fiedler, E.; Grossmann, G.; Kersebohm, D. B.; Weiss, G.; Witte, C., Methanol. Ullmann's Encyclopedia of Industrial Chemistry 2012.

42

63. Spath, P. L.; Dayton, D. C. Preliminary screening-technical and economic assessment of synthesis gas to fuels and chemicals with emphasis on the potential for biomass-derived syngas; DTIC Document: 2003. 64. LeBlanc, J.; Schneider, R.; Strait, R. B., Production of methanol. CHEMICAL INDUSTRIESNEW YORK-MARCEL DEKKER- 1994, 51-51. 65. Rahimpour, M. R.; Khosravanipour Mostafazadeh, A.; Barmaki, M. M., Application of hydrogen-permselective Pd-based membrane in an industrial single-type methanol reactor in the presence of catalyst deactivation. Fuel Processing Technology 2008, 89, (12), 13961408. 66. Borgwardt, R. H., Methanol production from biomass and natural gas as transportation fuel. Industrial & engineering chemistry research 1998, 37, (9), 3760-3767. 67. Heydorn, E.; Diamond, B.; Lilly, R.; Kornosky, R., Commercial-scale demonstration of the liquid phase methanol (LPMEOH) process. Prepared for the US DOE, National Energy Technology Laboratory 2003. 68. Assessment, A. Commercial-Scale Demonstration of the Liquid Phase Methanol (LPMEOH™) Process; DOE/NETL‐2004/1199: 2003. 69. Rauch, R., Biomass gasification to produce synthesis gas for fuel cells, liquid fuels and chemicals. IEA Bioenergy Agreement, Task 2002, 33, 2001-2003. 70. Saito, M.; Takeuchi, M.; Fujitani, T.; Toyir, J.; Luo, S.; Wu, J.; Mabuse, H.; Ushikoshi, K.; Mori, K.; Watanabe, T., Advances in joint research between NIRE and RITE for developing a novel technology for methanol synthesis from CO2 and H2. Applied organometallic chemistry 2000, 14, (12), 763-772. 71. Specht, M.; Bandi, A.; Baumgart, F.; Murray, C.; Gretz, J., Synthesis of methanol from biomass/CO2 resources. Synthesis 1999, 2, 68.7. 72. Fletcher, J.; Roan, V.; Betts, D. In An Investigation of the Feasibility of Coal-Based Methanol for Application in Transportation Fuel Cell Systems, 3rd International Energy Conversion Engineering Conference, 2005; 2005; p 5606. 73. Marlière, P., Method for the enzymatic production of 3-hydroxy-3-methylbutyric acid from acetone and acetyl-coa. In Google Patents: 2015. 74. Marliere, P., Production of alkenes hydroxyalkanoic acids. In Google Patents: 2009.

by

enzymatic

decarboxylation

of

3-

75. Marliere, P., Method for producing an alkene comprising the step of converting an alcohol by an enzymatic dehydration step. Google Patents 2011. 76. Fukuda, H.; Ogawa, T.; Fujii, T., Method for producing hydrocarbon mixtures. In Google Patents: 1987. 77. Fukuda, H.; Fujii, T.; Sukita, E.; Tazaki, M.; Nagahama, S.; Ogawa, T., Reconstitution of the isobutene-forming reaction catalyzed by cytochrome P450 and P450 reductase from Rhodotorula minuta: decarboxylation with the formation of isobutene. Biochemical and biophysical research communications 1994, 201, (2), 516-522. 43

78. Fukuda, H.; Fujii, T.; Ogawa, T., Production of isobutene by Rhodotorula yeasts. Agricultural and Biological Chemistry 1985, 49, (5), 1541-1543. 79. Schäfer, F.; Muzica, L.; Schuster, J.; Treuter, N.; Rosell, M.; Harms, H.; Müller, R. H.; Rohwerder, T., Alkene formation from tertiary alkyl ether and alcohol degradation by Aquincola tertiaricarbonis L108 and Methylibium spp. Applied and Environmental Microbiology 2011, AEM. 00093-11. 80. Jhung, S. H.; Chang, J.-S., Trimerization of isobutene over solid acid catalysts. Catalysis surveys from Asia 2009, 13, (4), 229-236. 81. Fukuda, H.; Fujii, T.; Ogawa, T., Microbial production of C3-and C4-hydrocarbons under aerobic conditions. Agricultural and biological chemistry 1984, 48, (6), 1679-1682. 82. Burnes, E.; Wichelns, D.; Hagen, J. W., Economic and policy implications of public support for ethanol production in California's San Joaquin Valley. Energy Policy 2005, 33, (9), 1155-1167. 83. Marchionna, M.; Di Girolamo, M.; Patrini, R., Light olefins dimerization to high quality gasoline components. Catalysis Today 2001, 65, (2), 397-403. 84. Piera, E.; Téllez, C.; Coronas, J. n.; Menéndez, M.; SaIntamarı́a, J., Use of zeolite membrane reactors for selectivity enhancement: application to the liquid-phase oligomerization of i-butene. Catalysis today 2001, 67, (1), 127-138. 85. Parkinson, G., All sides pumped up for MTBE ban. Chemical Engineering 1999, 106, (6), 49. 86. Williams, P. R.; Cushing, C. A.; Sheehan, P. J., Data available for evaluating the risks and benefits of MTBE and ethanol as alternative fuel oxygenates. Risk Analysis 2003, 23, (5), 1085-1115. 87. Giwa, A., Sensitivity analysis of ETBE production process using Aspen PLUS. Sensitivity Analysis of ETBE Production Process Using Aspen PLUS 2013, 3, (1), 293-303. 88. Tau, L.-M.; Davis, B. H., Acid catalyzed formation of ethyl tertiary butyl ether (ETBE). Applied catalysis 1989, 53, (2-3), 263-271. 89. Larsen, G.; Lotero, E.; Marquez, M.; Silva, H., Ethyl tert-butyl ether (ETBE) synthesis on H-mordenite: Gas-phase kinetics and DRIFTS studies. Journal of Catalysis 1995, 157, (2), 645655. 90. Fite, C.; Iborra, M.; Tejero, J.; Izquierdo, J. F.; Cunill, F., Kinetics of the liquid-phase synthesis of ethyl tert-butyl ether (ETBE). Industrial & engineering chemistry research 1994, 33, (3), 581-591. 91. Ancillotti, F.; Mauri, M. M.; Pescarollo, E., Ion exchange resin catalyzed addition of alcohols to olefins. Journal of Catalysis 1977, 46, (1), 49-57. 92. Kaitale, T.; Merikallio, E.; Rautianen, E. In The Application of Higher Ethers as Gasoline Components, 2nd EFOA Conference, Rome, Italy, 1987; 1987.

44

93. Gicquel, A.; Torck, B., Synthesis of methyl tertiary butyl ether catalyzed by ionexchange resin. Influence of methanol concentration and temperature. Journal of Catalysis 1983, 83, (1), 9-18. 94. Tong, L.; Chen, L.; Ye, Y.; Qi, Z., Kinetic studies on the dimerization of isobutene with Ni/Al 2 O 3 as a catalyst for reactive distillation process. Chinese Journal of Chemical Engineering 2015, 23, (3), 520-527. 95. Izquierdo, J.; Vila, M.; Tejero, J.; Cunill, F.; Iborra, M., Kinetic study of isobutene dimerization catalyzed by a macroporous sulphonic acid resin. Applied Catalysis A: General 1993, 106, (1), 155-165. 96. Yaocíhuatl, M. G.; Martín, H. L.; Jorge, A. C., Dimerization of isobutene over nickel modified zeolites to obtain isooctene. Catalysis letters 2006, 110, (1-2), 107-113. 97. Yoon, J. W.; Chang, J.-S.; Lee, H.-D.; Kim, T.-J.; Jhung, S. H., Trimerization of isobutene over a zeolite beta catalyst. Journal of Catalysis 2007, 245, (1), 253-256. 98. Yoon, J. W.; Lee, J. H.; Chang, J.-S.; Choo, D. H.; Lee, S. J.; Jhung, S. H., Trimerization of isobutene over zeolite catalysts: Remarkable performance over a ferrierite zeolite. Catalysis Communications 2007, 8, (6), 967-970. 99. Behr, A.; Rentmeister, N.; Seidensticker, T.; Vosberg, J.; Peitz, S.; Maschmeyer, D., Highly Selective Dimerization and Trimerization of Isobutene to Linearly Linked Products by Using Nickel Catalysts. Chemistry–An Asian Journal 2014, 9, (2), 596-601. 100. Honkela, M., Dimerisation of isobutene on acidic ion-exchange resins. Helsinki University of Technology: 2005. 101. Alcántara, R.; Alcántara, E.; Canoira, L.; Franco, M. a. J.; Herrera, M.; Navarro, A., Trimerization of isobutene over Amberlyst-15 catalyst. Reactive and functional polymers 2000, 45, (1), 19-27. 102. Grover, G.; Chaudhari, R., Selective dimerization of isobutene using a cation exchange resin catalyst. Recent Developments in catalysis-Theory and Practice, Narosa Publishing House, New Delhi 1991, 345. 103. Lee, J. S.; Yoon, J. W.; Halligudi, S. B.; Chang, J.-S.; Jhung, S. H., Trimerization of isobutene over WO x/ZrO 2 catalysts. Applied Catalysis A: General 2009, 366, (2), 299-303. 104. Mantilla, A.; Tzompantzi, F.; Ferrat, G.; Lopez-Ortega, A.; Alfaro, S.; Gomez, R.; Torres, M., Oligomerization of isobutene on sulfated titania: Effect of reaction conditions on selectivity. Catalysis today 2005, 107, 707-712. 105. Yoon, J. W.; Chang, J.-S.; Lee, H.-D.; Kim, T.-J.; Jhung, S. H., Trimerization of isobutene over cation exchange resins: Effect of physical properties of the resins and reaction conditions. Journal of Molecular Catalysis A: Chemical 2006, 260, (1), 181-186. 106. Yoon, J.-W.; Jhung, S.-H.; Chang, J.-S., Trimerization of isobutene over solid acid catalysts: comparison between cation-exchange resin and zeolite catalysts. Bulletin of the Korean Chemical Society 2008, 29, (2), 339-341.

45

107. Yoon, J.-W.; Jhung, S.-H.; Kim, T.-J.; Lee, H.-D.; Jang, N.-H.; Chang, J.-S., Trimerization of isobutene over solid acid catalysts under wide reaction conditions. Bulletin of the Korean Chemical Society 2007, 28, (11), 2075-2078. 108. Zhang, J.; Kanno, M.; Zhang, J.; Ohnishi, R.; Toriyabe, K.; Matsuhashi, H.; Kamiya, Y., Preferential oligomerization of isobutene in a mixture of isobutene and 1-butene over sodium-modified 12-tungstosilicic acid supported on silica. Journal of Molecular Catalysis A: Chemical 2010, 326, (1), 107-112. 109. Hauge, K.; Bergene, E.; Chen, D.; Fredriksen, G. R.; Holmen, A., Oligomerization of isobutene over solid acid catalysts. Catalysis today 2005, 100, (3), 463-466. 110. Di Girolamo, M.; Marchionna, M., Acidic and basic ion exchange resins for industrial applications. Journal of molecular catalysis A: chemical 2001, 177, (1), 33-40. 111. Mantilla, A.; Ferrat, G.; López-Ortega, A.; Romero, E.; Tzompantzi, F.; Torres, M.; OrtízIslas, E.; Gómez, R., Catalytic behavior of sulfated TiO 2 in light olefins oligomerization. Journal of Molecular Catalysis A: Chemical 2005, 228, (1), 333-338. 112. Tzompantzi, F.; Valverde-Herrera, M.; Rodríguez-González, J.; Gómez, R.; Mantilla, A.; Fernández, J. L.; Ortiz, H., Improved selectivity to C8-Olefins for isobutene oligomerization on NiO-W2O3/Al2O3 catalysts. Chemical Engineering Communications 2009, 196, (10), 11981205. 113. Zhang, J.; Ohnishi, R.; Okuhara, T.; Kamiya, Y., Preferential oligomerization of isobutene in mixtures of isobutene and 1-butene over 12-tungstosilicic acid supported on silica. Applied Catalysis A: General 2009, 353, (1), 68-73. 114. Lovallo, M. C.; Tsapatsis, M., Preferentially oriented submicron silicalite membranes. AIChE journal 1996, 42, (11), 3020-3029. 115. Jia, M.-D.; Chen, B.; Noble, R. D.; Falconer, J. L., Ceramic-zeolite composite membranes and their application for separation of vapor/gas mixtures. Journal of membrane science 1994, 90, (1), 1-10. 116. Sarkar, A.; Seth, D.; Ng, F. T.; Rempel, G. L., Selective Oligomerization of Isobutene on Lewis Acid Catalyst: Kinetic Modeling. Industrial & Engineering Chemistry Research 2014, 53, (49), 18982-18992. 117. Mantilla, A.; Tzompantzi, F.; Morales-Mendoza, G.; Jácome-Acatitla, G.; Gómez, R., Al2O3-Phosphated Green Catalysts with High Selectivity to Ecological Gasoline C= 8 by Dimerization of Isobutene. Journal of Nanoscience and Nanotechnology 2015, 15, (9), 72547257. 118. Huang, Y.-J.; Sun, Y.-M.; Chang, F.-L.; Hung, A. Y.; Wang, C.-T., The Feasibility Study of the Octane-Enhancing Process: Isobutylene Dimerization Using C4 Mixture. Journal of Environmental Science and Health, Part A 2003, 38, (11), 2661-2678. 119. Goortani, B. M.; Gaurav, A.; Deshpande, A.; Ng, F. T.; Rempel, G. L., Production of Isooctane from Isobutene: Energy Integration and Carbon Dioxide Abatement via Catalytic Distillation. Industrial & Engineering Chemistry Research 2015, 54, (14), 3570-3581.

46

120. Fritsch, D.; Randjelovic, I.; Keil, F., Application of a forced-flow catalytic membrane reactor for the dimerisation of isobutene. Catalysis today 2004, 98, (1), 295-308. 121. Ouni, T.; Honkela, M.; Kolah, A.; Aittamaa, J., Isobutene dimerisation in a miniplantscale reactor. Chemical Engineering and Processing: Process Intensification 2006, 45, (5), 329-339. 122. Kamath, R. S.; Qi, Z.; Sundmacher, K.; Aghalayam, P.; Mahajani, S. M., Process analysis for dimerization of isobutene by reactive distillation. Industrial & engineering chemistry research 2006, 45, (5), 1575-1582. 123. Kamath, R. S.; Qi, Z.; Sundmacher, K.; Aghalayam, P.; Mahajani, S. M., Comparison of reactive distillation with process alternatives for the isobutene dimerization reaction. Industrial & engineering chemistry research 2006, 45, (8), 2707-2714. 124. Honkela, M. L.; Krause, A. O. I., Kinetic modeling of the dimerization of isobutene. Industrial & engineering chemistry research 2004, 43, (13), 3251-3260. 125. Honkela, M.; Krause, A., Influence of polar components in the dimerization of isobutene. Catalysis letters 2003, 87, (3-4), 113-119. 126. Honkela, M. L.; Root, A.; Lindblad, M.; Krause, A. O. I., Comparison of ion-exchange resin catalysts in the dimerisation of isobutene. Applied Catalysis A: General 2005, 295, (2), 216-223. 127. O'Connor, C.; Kojima, M.; Schumann, W., The oligomerization of C 4 alkenes over cationic exchange resins. Applied catalysis 1985, 16, (2), 193-207. 128. Talwalkar, S.; Chauhan, M.; Aghalayam, P.; Qi, Z.; Sundmacher, K.; Mahajani, S., Kinetic studies on the dimerization of isobutene with ion-exchange resin in the presence of water as a selectivity enhancer. Industrial & engineering chemistry research 2006, 45, (4), 1312-1323. 129. Evans, T. I.; Karas, L. J.; Rameswaran, R., Selective olefin oligomerization. In Google Patents: 2002. 130. Di Girolamo, M.; Lami, M.; Marchionna, M.; Pescarollo, E.; Tagliabue, L.; Ancillotti, F., Liquid-phase etherification/dimerization of isobutene over sulfonic acid resins. Industrial & engineering chemistry research 1997, 36, (11), 4452-4458. 131. Talwalkar, S.; Mankar, S.; Katariya, A.; Aghalayam, P.; Ivanova, M.; Sundmacher, K.; Mahajani, S., Selectivity engineering with reactive distillation for dimerization of C4 olefins: experimental and theoretical studies. Industrial & engineering chemistry research 2007, 46, (10), 3024-3034. 132.

Bowman, W. G.; Stadig, W. P., Dimerization of isobutene. In Google Patents: 1978.

133. Honkela, M. L.; Krause, A. O. I., Influence of linear butenes in the dimerization of isobutene. Industrial & engineering chemistry research 2005, 44, (14), 5291-5297. 134. Dakka, J. M.; Geelen, M. O. J.; Mathys, G. M.; Allen, P. W., Process for the selective dimerization of isobutene. In Google Patents: 2005.

47

135. Antunes, B. M.; Rodrigues, A. E.; Lin, Z.; Portugal, I.; Silva, C. M., Alkenes oligomerization with resin catalysts. Fuel Processing Technology 2015, 138, 86-99. 136. Peters, M. W.; Taylor, J. D.; Jenni, M.; Manzer, L. E.; Henton, D. E., Integrated process to selectively convert renewable isobutanol to p-xylene. In Google Patents: 2010. 137. Bonilla, A.; Baudouin, D.; Pérez-Ramírez, J., Desilication of ferrierite zeolite for porosity generation and improved effectiveness in polyethylene pyrolysis. Journal of Catalysis 2009, 265, (2), 170-180. 138. Cadenas, M.; Bringué, R.; Fité, C.; Ramírez, E.; Cunill, F., Liquid-phase oligomerization of 1-hexene catalyzed by macroporous ion-exchange resins. Topics in Catalysis 2011, 54, (1315), 998-1008. 139. Chauvin, Y.; Olivier, H.; Wyrvalski, C. N.; Simon, L. C.; de Souza, R. F., Oligomerization ofn-Butenes Catalyzed by Nickel Complexes Dissolved in Organochloroaluminate Ionic Liquids. Journal of Catalysis 1997, 165, (2), 275-278. 140.

Kimura, N.; Morikita, T.; Hamamatsu, T., New Butene Dimerization Technology.

141. Cavani, F.; Girotti, G.; Terzoni, G., Effect of water in the performance of the “solid phosphoric acid” catalyst for alkylation of benzene to cumene and for oligomerization of propene. Applied Catalysis A: General 1993, 97, (2), 177-196. 142. Krawietz, T. R.; Lin, P.; Lotterhos, K. E.; Torres, P. D.; Barich, D. H.; Clearfield, A.; Haw, J. F., Solid phosphoric acid catalyst: A multinuclear NMR and theoretical study. Journal of the American Chemical Society 1998, 120, (33), 8502-8511. 143. Hunszinger, P.; Järvelin, H.; Purola, V.; Nurminen, M.; Khalil, M.; Birkhoff, R. In Startup and operation of the first on-purpose isooctane unit at Alberta Envirofuels Inc., Edmonton, Canada, NPRA 2003 Annual Meeting, San Antonio, TX, 2003; 2003. 144. Birkhoff, R.; Nurminen, M., NExOCTANE technology for isooctane production. Handbook of Petroleum Refining Processes 2004, 3. 145. Sahay, N.; McGuirk, T.; Marchionna, M.; Tagliabue, L. In Low cost conversion of MTBE units to alternative gasoline blending components production, National Petrochemical & Refiners Association Annual Meeting, Washington DC, 2002; 2002. 146. Sanfilippo, D., Dehydrogenation of paraffins; key technology for petrochemicals and fuels. Cattech 2000, 4, (1), 56-73. 147.

Venuto, P. B.; Habib Jr, E. T., Fluid catalytic cracking with zeolite catalysts. 1979.

148. Avidan, A. A., Fluid catalytic cracking. In Circulating fluidized beds, Springer: 1997; pp 466-488. 149. Zimmermann, H.; Walzl, R., Ethylene. Ullmann's Encyclopedia of Industrial Chemistry 2009. 150. Diercks, R.; Arndt, J. D.; Freyer, S.; Geier, R.; Machhammer, O.; Schwartze, J.; Volland, M., Raw material changes in the chemical industry. Chemical engineering & technology 2008, 31, (5), 631-637. 48

151. Cares, W. R.; Hightower, J. W., Ferrite spinels as catalysts in the oxidative dehydrogenation of butenes. Journal of Catalysis 1971, 23, (2), 193-203. 152. Toledo-Antonio, J.; Nava, N.; Martınez, M.; Bokhimi, X., Correlation between the magnetism of non-stoichiometric zinc ferrites and their catalytic activity for oxidative dehydrogenation of 1-butene. Applied Catalysis A: General 2002, 234, (1), 137-144. 153. Gibson, M. A.; Hightower, J. W., Oxidative dehydrogenation of butenes over magnesium ferrite: Catalyst deactivation studies. Journal of Catalysis 1976, 41, (3), 431-439. 154. Qiu, F.-Y.; Weng, L.-T.; Sham, E.; Ruiz, P.; Delmon, B., Effect of added Sb 2 O 4, BiPO 4 or SnO 2 on the catalytic properties of ZnFe 2 O 4 in the oxidative dehydrogenation of butene to butadiene. Applied catalysis 1989, 51, (1), 235-253. 155. Rennard, R.; Kehl, W., Oxidative dehydrogenation of butenes over ferrite catalysts. Journal of Catalysis 1971, 21, (3), 282-293. 156. Park, J.-H.; Noh, H.; Park, J. W.; Row, K. H.; Jung, K. D.; Shin, C.-H., Oxidative dehydrogenation of n-butenes to 1, 3-butadiene over BiMoFe0. 65P x catalysts: effect of phosphorous contents. Research on Chemical Intermediates 2011, 37, (9), 1125-1134. 157. Weissermel, K.; Arpe, H., Industrial Organic Chemistry, ; VCH: New York, 1993.(b) Aromatic Hydrogenation Catalysis: A Review. Stanislaus, A.; Cooper, BH. Catal. Rev.-Sci. Eng 1994, 36, 75. 158. Noll, H.; Bland, R.; Kelso, G., Effect of Temperature in TCC Catalytic Cracking. Houdry Pioneer 1947, 2, (2). 159. Greensfelder, B.; Voge, H., Catalytic Cracking of Pure Hydrocarbons. Industrial & Engineering Chemistry 1945, 37, (6), 514-520. 160.

Voge, H.; Good, M., BS Greensfelder Proc. 3rd World Pet. Congr 1951, 4, 124.

161. Greensfelder, B.; Voge, H.; Good, G., Catalytic and thermal cracking of pure hydrocarbons: Mechanisms of Reaction. Industrial & Engineering Chemistry 1949, 41, (11), 2573-2584. 162. Lylykangas, M.; Rautanen, P.; Krause, A., Liquid‐phase hydrogenation kinetics of isooctenes on Ni/Al2O3. AIChE journal 2003, 49, (6), 1508-1515. 163. Karinen, R. S.; Krause, A. O. I., Reactivity of some C 8-alkenes in etherification with methanol. Applied Catalysis A: General 1999, 188, (1), 247-256. 164. Karinen, R.; Linnekoski, J.; Krause, A., Etherification of C5-and C8-alkenes with C1-to C4-alcohols. Catalysis letters 2001, 76, (1-2), 81-87. 165. Klotz Rabello, C. R.; Wolff, M.; Silva, M.; Menegassi, R.; de Souza, G.; Hu, X. In PETROBRAS MTBE Unit Retrofit to Isoctane Production, 18th World Petroleum Congress, 2005; World Petroleum Congress: 2005. 166. Baerlocher, C.; McCusker, L. B.; Olson, D. H., Atlas of zeolite framework types. Elsevier: 2007.

49

167. Bakare, I. A.; Muraza, O.; Ganiyu, S. A.; Hakeem, A. S.; Yamani, Z. H.; Al-Amer, A. M. J., Synthesis of zeolite–magadiite composites: Effects of co-solvent and aluminum source. Particuology. 168. Vaughan, P. A., The crystal structure of the zeolite ferrierite. Acta Crystallographica 1966, 21, (6), 983-990. 169. Pantu, P.; Boekfa, B.; Sunpetch, B.; Limtrakul, J., Nanocavity effects on N2O decomposition on different types of Fe-Zeolites (Fe-FER, Fe-BEA, Fe-ZSM-5 and Fe-FAU): a combined theoretical and experimental study. Chemical Engineering Communications 2008, 195, (11), 1477-1485. 170. Mauvezin, M.; Delahay, G.; Kisslich, F.; Coq, B.; Kieger, S., Catalytic Reduction of N2O by NH3 in Presence of Oxygen using Fe‐exchanged Zeolites. Catalysis letters 1999, 62, (1), 41-44. 171. Aramendía, M. A.; Borau, V.; Jiménez, C.; Marinas, J. M.; Roldán, R.; Romero, F. J.; Urbano, F. J., Catalytic Application of Zeolites in the Methanol Conversion to Hydrocarbons. Chemistry Letters 2002, (7), 672-673. 172. Anderson, B. G.; Schumacher, R. R.; van Duren, R.; Singh, A. P.; van Santen, R. A., An attempt to predict the optimum zeolite-based catalyst for selective cracking of naphtharange hydrocarbons to light olefins. Journal of Molecular Catalysis A: Chemical 2002, 181, (1– 2), 291-301. 173. Kibby, C.; Perrotta, A.; Massoth, F., Composition and catalytic properties of synthetic ferrierite. Journal of Catalysis 1974, 35, (2), 256-272. 174. Jiménez, C.; Romero, F. J.; Roldán, R.; Marinas, J. M.; Gómez, J. P., Hydroisomerization of a hydrocarbon feed containing n-hexane, n-heptane and cyclohexane on zeolite-supported platinum catalysts. Applied Catalysis A: General 2003, 249, (1), 175-185. 175. Fottinger, K.; Kinger, G.; Vinek, H., 1-Pentene isomerization over FER and BEA. Applied Catalysis A: General 2003, 249, (2), 205-212. 176. Mooiweer, H. H.; de Jong, K. P.; Kraushaar-Czarnetzki, B.; Stork, W. H. J.; Krutzen, B. C. H., Skeletal isomerisation of olefins with the zeolite Ferrierite as catalyst. In Studies in Surface Science and Catalysis, J. Weitkamp, H. G. K. H. P.; Hölderich, W., Eds. Elsevier: 1994; Vol. Volume 84, pp 2327-2334. 177. Pellet, R.; Casey, D.; Huang, H.; Kessler, R.; Kuhlman, E.; Oyoung, C.; Sawicki, R.; Ugolini, J., Isomerization of n-butene to isobutene by ferrierite and modified ferrierite catalysts. Journal of Catalysis 1995, 157, (2), 423-435. 178. Guisnet, M.; Andy, P.; Gnep, N.; Travers, C.; Benazzi, E., Mechanisms of the skeletal isomerization of n-butene over a HFER zeolite. Influence of coke deposits. Studies in Surface Science and Catalysis 1997, 105, 1365-1372. 179. Guisnet, M.; Andy, P.; Boucheff, Y.; Gnep, N.; Travers, C.; Benazzi, E., Selective isomerization of n-butenes into isobutene over aged H-ferrierite catalyst: nature of the active species. Catalysis letters 1998, 50, (3-4), 159-164.

50

180. Guisnet, M.; Andy, P.; Gnep, N.; Benazzi, E.; Travers, C., Skeletal Isomerization ofnButenes: I. Mechanism ofn-Butene Transformation on a Nondeactivated H-Ferrierite Catalyst. Journal of Catalysis 1996, 158, (2), 551-560. 181. Oyoung, C.-L.; Pellet, R. J.; Casey, D. G.; Ugolini, J. R.; Sawicki, R. A., Skeletal isomerization of 1-butene on 10-member ring zeolite catalysts. Journal of Catalysis 1995, 151, (2), 467-469. 182. Cañizares, P.; Carrero, A., Dealumination of ferrierite by ammonium hexafluorosilicate treatment: characterization and testing in the skeletal isomerization of nbutene. Applied Catalysis A: General 2003, 248, (1), 227-237. 183. Canizares, P.; Carrero, A.; Sanchez, P., Isomerization of n-butene over ferrierite zeolite modified by silicon tetrachloride treatment. Applied Catalysis A: General 2000, 190, (1), 93-105. 184. Pazè, C.; Sazak, B.; Zecchina, A.; Dwyer, J., FTIR and UV-Vis spectroscopic study of interaction of 1-butene on H-ferrierite zeolite. The Journal of Physical Chemistry B 1999, 103, (45), 9978-9986. 185. Stepanov, A. G.; Luzgin, M. V.; Arzumanov, S. S.; Ernst, H.; Freude, D., n-Butene Conversion on H-Ferrierite Studied by 13 C MAS NMR. Journal of Catalysis 2002, 211, (1), 165-172. 186. Ivanov, P.; Papp, H., FT-IR Study of the Isomerization of n-Butene over Different Zeolites. Langmuir 2000, 16, (20), 7769-7772. 187. Čejka, J.; Wichterlová, B.; Sarv, P., Extent of monomolecular and bimolecular mechanism in n-butene skeletal isomerization to isobutene over molecular sieves. Applied Catalysis A: General 1999, 179, (1–2), 217-222. 188. Yang, B.; Jiang, J.-g.; Xu, H.; Ji, P.; Wu, P., Sub-zeolite of FER topology derived from an interlayer modification of PLS-3 lamellar precursor. Microporous and Mesoporous Materials 2015, 203, 54-62. 189. Xue, T.; Liu, H.; Wang, Y. M., Synthesis of hierarchical ferrierite using piperidine and tetramethylammonium hydroxide as cooperative structure-directing agents. RSC Advances 2015, 5, (16), 12131-12138. 190. Frontera, P.; Candamano, S.; Macario, A.; Crea, F.; Scarpino, L. A.; Antonucci, P. L., Ferrierite zeolitic thin-layer on cordierite honeycomb support by clear solutions. Materials Letters 2013, 104, 72-75. 191. Lee, Y.; Park, M. B.; Kim, P. S.; Vicente, A. l.; Fernandez, C.; Nam, I.-S.; Hong, S. B., Synthesis and Catalytic Behavior of Ferrierite Zeolite Nanoneedles. ACS Catalysis 2013, 3, (4), 617-621. 192. Zhao, Z.; Zhang, W.; Ren, P.; Han, X.; Müller, U.; Yilmaz, B.; Feyen, M.; Gies, H.; Xiao, F.S.; De Vos, D., Insights into the topotactic conversion process from layered silicate RUB-36 to FER-type zeolite by layer reassembly. Chemistry of Materials 2013, 25, (6), 840-847.

51

193. Pinar, A. B.; Gómez-Hortigüela, L.; McCusker, L. B.; Pérez-Pariente, J., Controlling the Aluminum Distribution in the Zeolite Ferrierite via the Organic Structure Directing Agent. Chemistry of Materials 2013, 25, (18), 3654-3661. 194. Khitev, Y. P.; Kolyagin, Y. G.; Ivanova, I.; Ponomareva, O.; Thibault-Starzyk, F.; Gilson, J.-P.; Fernandez, C.; Fajula, F., Synthesis and catalytic properties of hierarchical micro/mesoporous materials based on FER zeolite. Microporous and Mesoporous Materials 2011, 146, (1), 201-207. 195. Chauhan, N. L.; Murthy, Z.; Das, J.; Parikh, P. A., Synthesis of zeolite ferrierite-Role of emulsifiers. Indian Journal of Chemical Technology 2011, 18, (5). 196. Verboekend, D.; Caicedo-Realpe, R.; Bonilla, A.; Santiago, M.; Pérez-Ramírez, J., Properties and functions of hierarchical ferrierite zeolites obtained by sequential postsynthesis treatments. Chemistry of Materials 2010, 22, (16), 4679-4689. 197. Ruan, J.; Wu, P.; Slater, B.; Zhao, Z.; Wu, L.; Terasaki, O., Structural characterization of interlayer expanded zeolite prepared from ferrierite lamellar precursor. Chemistry of Materials 2009, 21, (13), 2904-2911. 198. Cheng, X.-w.; Wang, J.; Guo, J.; He, H.-y.; Long, Y.-c., FER zeolite crystallized in THF–Na 2 O–SiO 2–Al 2 O 3–H 2 O reactant system containing catalytic amount of organic additives. Microporous and Mesoporous Materials 2009, 119, (1), 60-67. 199. Wuamprakhon, P.; Wattanakit, C.; Warakulwit, C.; Yutthalekha, T.; Wannapakdee, W.; Ittisanronnachai, S.; Limtrakul, J., Direct synthesis of hierarchical ferrierite nanosheet assemblies via an organosilane template approach and determination of their catalytic activity. Microporous and Mesoporous Materials 2016, 219, 1-9. 200. Almeida, R. K.; Gómez-Hortigüela, L.; Pinar, A. B.; Peréz-Pariente, J., Synthesis of ferrierite by a new combination of co-structure-directing agents: 1, 6-bis (Nmethylpyrrolidinium) hexane and tetramethylammonium. Microporous and Mesoporous Materials 2016, 232, 218-226. 201. Kumar, P.; Sung, C.-Y.; Muraza, O.; Cococcioni, M.; Al Hashimi, S.; McCormick, A.; Tsapatsis, M., H2S adsorption by Ag and Cu ion exchanged faujasites. Microporous and Mesoporous Materials 2011, 146, (1–3), 127-133. 202. Muraza, O.; Rebrov, E. V.; Chen, J.; Putkonen, M.; Niinistö, L.; De Croon, M. H.; Schouten, J. C., Microwave-assisted hydrothermal synthesis of zeolite Beta coatings on ALD-modified borosilicate glass for application in microstructured reactors. Chemical Engineering Journal 2008, 135, S117-S120. 203. Lin, Z. S.; Huang, Y., Syntheses of High-Silica Zeolites in Urea/Choline Chloride Deep Eutectic Solvent. Canadian Journal of Chemistry 2016, (ja). 204. Baran, R.; Valentin, L.; Dzwigaj, S., Incorporation of Mn into the vacant T-atom sites of a BEA zeolite as isolated, mononuclear Mn: FTIR, XPS, EPR and DR UV-Vis studies. Physical Chemistry Chemical Physics 2016, 18, (17), 12050-12057.

52

205. Nasser, G. A.; Kurniawan, T.; Tago, T.; Bakare, I. A.; Taniguchi, T.; Nakasaka, Y.; Masuda, T.; Muraza, O., Cracking of n-hexane over hierarchical MOR zeolites derived from natural minerals. Journal of the Taiwan Institute of Chemical Engineers 2016, 61, 20-25. 206. Khalil, U.; Muraza, O., Microwave-assisted hydrothermal synthesis of mordenite zeolite: Optimization of synthesis parameters. Microporous and Mesoporous Materials 2016. 207. Li, L.; Yang, J.; Li, J.; Han, P.; Wang, J.; Zhao, Y.; Wang, J.; Lu, J.; Yin, D.; Zhang, Y., Synthesis of high performance mordenite membranes from fluoride-containing dilute solution under microwave-assisted heating. Journal of Membrane Science 2016, 512, 83-92. 208. AONO, H.; KAJI, N.; ITAGAKI, Y.; JOHAN, E.; MATSUE, N., Synthesis of mordenite and its composite material using chemical reagents for Cs decontamination. Journal of the Ceramic Society of Japan 2016, 124, (5), 617-623. 209. Ahmadpour, J.; Taghizadeh, M., One-Pot Synthesis of Hierarchically Mesoporous ZSM5 Using Different Combinations of Mesogenous Templates. Synthesis and Reactivity in Inorganic, Metal-Organic, and Nano-Metal Chemistry 2016, 46, (8), 1133-1141. 210. Rilyanti, M.; Mukti, R. R.; Kadja, G. T.; Ogura, M.; Nur, H.; Ng, E.-P., On the drastic reduction of organic structure directing agent in the steam-assisted crystallization of zeolite with hierarchical porosity. Microporous and Mesoporous Materials 2016, 230, 30-38. 211. Han, S.; Wang, Z.; Meng, L.; Jiang, N., Synthesis of uniform mesoporous ZSM-5 using hydrophilic carbon as a hard template. Materials Chemistry and Physics 2016, 177, 112-117. 212. Pashkova, V.; Tokarova, V.; Brabec, L.; Dedecek, J., Self-templating synthesis of hollow spheres of zeolite ZSM-5 from spray-dried aluminosilicate precursor. Microporous and Mesoporous Materials 2016, 228, 59-63. 213. Zhang, B.; Li, X.; Wu, Q.; Zhang, C.; Yu, Y.; Lan, M.; Wei, X.; Ying, Z.; Liu, T.; Liang, G., Synthesis of Ni/mesoporous ZSM-5 for direct catalytic conversion of cellulose to hexitols: modulating the pore structure and acidic sites via a nanocrystalline cellulose template. Green Chemistry 2016, 18, (11), 3315-3323. 214. Zhang, H.; Wang, L.; Zhang, D.; Meng, X.; Xiao, F.-S., Mesoporous and Al-rich MFI crystals assembled with aligned nanorods in the absence of organic templates. Microporous and Mesoporous Materials 2016. 215. Cundy, C. S.; Cox, P. A., The hydrothermal synthesis of zeolites: history and development from the earliest days to the present time. Chemical Reviews 2003, 103, (3), 663-702. 216. Choi, M.; Na, K.; Kim, J.; Sakamoto, Y.; Terasaki, O.; Ryoo, R., Stable single-unit-cell nanosheets of zeolite MFI as active and long-lived catalysts. Nature 2009, 461, (7261), 246249. 217. Ng, F.; Creaser, D., Ethylene dimerization over modified nickel exchanged Y-zeolite. Applied Catalysis A: General 1994, 119, (2), 327-339. 218. Torres, M.; Gutiérrez, M.; Mugica, V.; Romero, M.; López, L., Oligomerization of isobutene with a beta-zeolite membrane: Effect of the acid properties of the catalytic membrane. Catalysis today 2011, 166, (1), 205-208. 53

219. Park, D. H.; Kim, S.-S.; Pinnavaia, T. J.; Tzompantzi, F.; Prince, J.; Valente, J. S., Selective isobutene oligomerization by mesoporous MSU-SBEA catalysts. The Journal of Physical Chemistry C 2011, 115, (13), 5809-5816. 220. Sloan, H.; Birkhoff, R.; Gilbert, M.; Nurminen, M.; Pyhälahti, A., ISOOCTANE PRODUCTION FROM C~ 4'S AS AN ALTERNATIVE TO MTBE. NATIONAL PETROCHEMICAL AND REFINERS ASSOCIATION-PUBLICATIONS-ALL SERIES 2000. 221. Trotta, R.; Marchionna, M.; Amoretto, F., An iso-octane technology to produce alkylate streams. Today’s Refinery 1999, 10, 21-24. 222. Sarkar, A., A catalytic distillation process for one-step production of isooctane from isobutene-process development, modelling and analysis. Chemical Engineering2005, University of Waterloo: Waterloo 2005, 364. 223.

Ng, F. T.; Rempel, G. L., Catalytic distillation. Encyclopedia of Catalysis 2003.

224. Karinen, R. S.; Lylykangas, M. S.; Krause, A. O. I., Reaction equilibrium in the isomerization of 2, 4, 4-trimethyl pentenes. Industrial & engineering chemistry research 2001, 40, (4), 1011-1015. 225. Karinen, R. S.; Krause, A. O. I., Kinetic Model for the Etherification of 2, 4, 4-Trimethyl1-pentene and 2, 4, 4-Trimethyl-2-pentene with Methanol. Industrial & engineering chemistry research 2001, 40, (26), 6073-6080. 226. Zhang, X.; Zhong, J.; Wang, J.; Zhang, L.; Gao, J.; Liu, A., Catalytic performance and characterization of Ni-doped HZSM-5 catalysts for selective trimerization of n-butene. Fuel Processing Technology 2009, 90, (7), 863-870. 227. Brown, T. C.; Miron, D. J.; Alanazi, A. K.; Le Minh, C., Rate Parameter Distributions for Isobutane Dehydrogenation and Isobutene Dimerization and Desorption over HZSM-5. Catalysts 2013, 3, (4), 922-941. 228. Koskinen, R.; Turunen, H.; Tiitta, M.; Keiski, R. L., Extended activity of zeolite catalysts with CO 2 as reaction medium. Chemical engineering journal 2012, 213, 235-244. 229. Bakare, I. A.; Muraza, O.; Al-Amer, A. M.; Yamani, Z. H., The effect of non-ionic surfactant in the microwave-assisted synthesis of MTT zeolite optimized by Taguchi method. Journal of the Taiwan Institute of Chemical Engineers 2015, 50, 314-321. 230. 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. Chemical Engineering Journal 2013, 226, 367-376. 231. 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. 232. Bakare, I. A.; Muraza, O.; Taniguchi, T.; Tago, T.; Nasser, G.; Yamani, Z. H.; Masuda, T., Steam-assisted catalytic cracking of n-hexane over La-Modified MTT zeolite for selective propylene production. Journal of Analytical and Applied Pyrolysis 2015, 116, 272-280.

54

233. Bakare, I. A.; Muraza, O.; Kurniawan, T.; Yamani, Z. H.; Shafei, E. N.; Punetha, A. K.; Choi, K.-H.; Yokoi, T., Hydrothermal stability of MTT zeolite in hot water: The role of La and Ce. Microporous and Mesoporous Materials. 234. Yoon, J. W.; Lee, J. S.; Jhung, S. H.; Lee, K.-Y.; Chang, J.-S., Oligomerization of isobutene over aluminum chloride-loaded USY zeolite catalysts. Journal of Porous Materials 2009, 16, (6), 631-634. 235. Yoon, J. W.; Jhung, S. H.; Choo, D. H.; Lee, S. J.; Lee, K.-Y.; Chang, J.-S., Oligomerization of isobutene over dealuminated Y zeolite catalysts. Applied Catalysis A: General 2008, 337, (1), 73-77. 236. Ngandjui, L. T.; Thyrion, F., Kinetic study of the oligomerization of isobutene over Hmordenite. Chemical Engineering and Processing: Process Intensification 1992, 31, (1), 1-6. 237. Verboekend, D.; Nuttens, N.; Locus, R.; Van Aelst, J.; Verolme, P.; Groen, J.; PérezRamírez, J.; Sels, B., Synthesis, characterisation, and catalytic evaluation of hierarchical faujasite zeolites: milestones, challenges, and future directions. Chemical Society Reviews 2016. 238. Zhao, J.; Yin, Y.; Li, Y.; Chen, W.; Liu, B., Synthesis and characterization of mesoporous zeolite Y by using block copolymers as templates. Chemical Engineering Journal 2016, 284, 405-411. 239. Yuan, D.; He, D.; Xu, S.; Song, Z.; Zhang, M.; Wei, Y.; He, Y.; Xu, S.; Liu, Z.; Xu, Y., Imidazolium-based ionic liquids as novel organic SDA to synthesize high-silica Y zeolite. Microporous and Mesoporous Materials 2015, 204, 1-7.

55

30

28

Million tons per year (MTPY)

26

24

22

20

18

16

14

12

10 1992

1993

1994

1995

1997

1998

2000

2001

2002

2003

2004

Year Figure 1. MTBE capacity and demand in the last decades [10–15].

2005

2011

Fluid catalytic cracking (FCC)

C4 olefins

Naphtha and Ethane steam cracking (SC)

C4 olefins

Oxidative dehydrogenation (ODH)

Dimerization

isobutylene

Skeletal isomerization of n-butenes

C4 olefins iso-octene

H2

Hydrogenation

iso-octane

As an octane booster

Figure 2. Flowscheme isooctane production from isobutylene using NExOCTANE technology (adapted from KBR technology).

CH3

H3C C

CH2

+

C

H2C

H3C

Isobutylene

Catalyst

CH3

H

H3C C

CH2

C

CH3

Isobutylene H

C

CH3

C

CH3

Catalyst

CH3

H3C

CH3 C

CH2

+

CH3 H C

CH3

CH3 H

C C H2

CH3

C

Catalyst

CH3

C

HC

H3C

Isooctene (2,4,4-trimethyl-2-pentene)

H3C

C

H3C

CH3 H H3C

CH3

HC

+

H

H3C

C

CH3

C

C CH2

CH3 H

Isooctene (2,4,4-trimethyl-1-pentene) CH3 H H3C

C

CH3

C

C CH2

CH3 H

CH3

CH3 H H3C

C

C

CH3

C

CH3

+ H

H

H2 (Hydrogenation)

H3C

CH3

C CH3

H3C

C

C

CH3 H

CH3

CH3 +

C

H

H

H2 (Hydrogenation)

CH2

Isooctene (2,4,4-trimethyl-1-pentene)

CH3

C

C

CH3

H

H

Isooctane

Isooctene (2,4,4-trimethyl-2-pentene) CH3 H

H

H3C

C CH3

H

CH3

C

C

H

H

CH3

Isooctane

Figure 3. Mechanism of isooctane production from dimerization of isobutylene followed by hydrogenation.

Novel organic template

Shorten synthesis time Al58Si134 O384 Low Lewis/Brønsted acid site ratio

USY zeolite

Figure 4. Strategies to improve USY zeolite for dimerization of isobutylene to isooctane.

C4 olefins

Dimerization

Isobutylene Organic template

Iso-octene

Shorten synthesis time Al58Si134 O384

USY zeolite H2

Hydrogenation

Iso-octane

Low Lewis/Brønsted acid sites ratio