Vegetable oil to Biolubricants: Review on Advanced Porous Catalysts

3 days ago - Ferrierite can be produced using low-cost OSDA. The possibility to design two-dimensional pore zeolites with pore mouth selectivity was a...
0 downloads 0 Views 1MB Size
Subscriber access provided by University of South Dakota

Review

Vegetable oil to Biolubricants: Review on Advanced Porous Catalysts Ahmad Masudi, and Oki Muraza Energy Fuels, Just Accepted Manuscript • Publication Date (Web): 29 Aug 2018 Downloaded from http://pubs.acs.org on August 30, 2018

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 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 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.

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.

Page 1 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

Vegetable oil to Biolubricants: Review on Advanced Porous Catalysts

Ahmad Masudi1, Oki Muraza2* 1

Department of Chemical Process Engineering, Malaysia-Japan International Institute of Technology (MJIIT)

Universiti Teknologi Malaysia Kuala Lumpur, Jalan Sultan Yahya Petra, 54100 Kuala Lumpur, Malaysia 2

Center of Research Excellence in Nanotechnology and Chemical Engineering Department King Fahd University of Petroleum & Minerals, Dhahran 31261, Saudi Arabia

Abstract Vegetable oil is one of the most potential sustainable feedstocks to produce fuels and chemicals. The article emphasizes on isomerization of fatty acid as an important path for biolubricant production. The role of solid acid catalysts including zeolites was highlighted to design better isomerization catalysts. The isomerization is favored mesoporous site with intermediate Brønsted acid strength which is also enhanced after metal doping on porous surface. Hierarchical ferrierite (FER) catalyst showed a best selective isomerization with the 10-MR cavities, which can be regenerated easily. Ferrierite can be produced using low-cost OSDA. The possibility to design two-dimensional pore zeolites with pore mouth selectivity was also discussed. Moreover, the challenges for biolubricant formulation with focus on palm oil were also discussed with detailed comparison to other vegetable oils. The highest palm oil conversion was achieved over based catalyst namely Sr doped calcium oxide with low catalyst dosage. However, the biolubricant based palm oil still needs many advancements to achieve industrial standards. Keywords: Isomerization; fatty acids; Brønsted acid; ferrierite; biolubricants. *Corresponding author: [email protected]

1 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 45

1. Introduction The rising of global concern to climate change, made the vegetable oil become an alternative option for future fuels and oleo-chemical demands. Among the resources, palm oil is the largest contributor with 35 wt.% or 64 million tons of bio-oil consumption in 2014. The forecast utilization of palm oil would rise to 78 million in 2020. This rapid growth of palm oil consumption is due to its highest ratio of yield to price than other vegetable oil such as soybean and sunflower. The oil yield of palm, soybean and sunflower namely 4, 1 and 0.37 MT/ha, respectively 1-3. The largest world palm oil producer is Indonesia, which is followed by Malaysia and Thailand. Both Indonesia and Malaysia were concerned to export biodiesel derived from palm oil. Meanwhile, Thailand focused for domestic consumption. The growth of Indonesian biodiesel in a year at 2010 was about 117% which was followed by land clearing of 20 milion hectares 4. On contrary, Thailand prevent export for pure biodiesel and allocated entire production for national demand

5

. In Malaysia, palm oil is among the largest national revenue after

petrochemical product

6-8

. A new palm oil plantation would lead to deforestation, Malaysian

government committed to replace its risk with plantation of cocoa, rubber and coconut 9. Palm oil is originated from mesocarp of palm with 45-50 wt.% of mesocarp. Extracted palm oil consists of triglyceride and free fatty acids (FFA). Free fatty acid is linear carboxyclic acid which contains of 12-24 of carbon. The types of FFA in palm oil are mainly consist of oleic acid (C18:1) and palmitic acid (C16:0). Although palmitic acid is not the main component in palm oil, its high content become the uniqueness of palm oil as compared with other vegetable oils 10. Fatty acids in vegetable oil are major green source to produce value added chemical such as surfactant, lubricant, soap and detergent. Global demand of this fatty acid was around 6

2 ACS Paragon Plus Environment

Page 3 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

million tons in 2011 and estimated to grow to 7.4 millions in 2016. The application of fatty acid could increase GDP of its city and province

11

. As example, coconut oil was transformed to jet

fuel over Mo-Ni/ߛ-Al2O3 in the presence of low hydrogen pressure. The product distribution could was varied according to time contact between coconut oil and catalysts. In the first 20 min, the catalytic system favored hydrodeoxygenation followed with cracking of ester carbonyl after 1 h

12

. In addition, polyunsaturated vegetable oil could be converted to conjugated oil over

Ru/USY catalyst. The key role of the fabrication was zeolite mesoporosity to facilitate better diffusion and elimination of reactive conjugated triglycerides 13. Saturated carbon chain in fatty acid is stable, while unsaturated carbon is unstable and prone to oxidation. Free fatty acid in vegetable oil consist of both saturated and saturated chain. The presence of branch in fatty acid was confirmed to increase cold flow properties, which is crucial to diverse biodiesel market range and increase efficiency. There are two methods to induce branch formation namely (i) direct isomerization from saturated and (ii) isomerization from unsaturated which is followed by hydrogenation. The research verified that the second method showed better conversion and higher selectivity as reported in many patents

14-17

. In

recent studies, it was concluded that increase of fatty acid feed decreased isomerization rate with constant activation energy

18

. In addition, isomerization reaction also competes with

oligomerization due to the same carbocation intermediates. Therefore, it was essential to develop catalyst with better isomerization conversion and selectivity. Among the catalysts, modified ߛAl2O3 with sulphate acid was successful for isomerization of 1-decanol to 1-decene

19

.

Meanwhile, NiW/SiO2-Al2O3 was effective for isomerization of C18 fatty acids at low WHSV (weight hour space velocity) 20. Pt/AlSBA-15 was reported as the best catalyst for isomerization of triglyceride at 345 °C and 50 bar. The product variations of triglyceride were fuels (biodiesel,

3 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

bioethanol, biogas oil, and bio-hydrogen)

21

Page 4 of 45

and branch paraffins which are potential as

biolubricant 22. Biolubricant is one of the most potential applications of branched fatty acid. Biolubricant could be prepared with esterification of free fatty acids, which is usually obtained over alkaline catalyst in industrial scale

23

. Generally, biolubricant which is prepared for certain application,

will not be suitable for other applications. Biolubricant needs many modifications to meet the standard due to different requirements in installed equipments. Some characteristics such as pour point, viscosity, flash point, viscosity index and oxidative stability are essential for its appropriate applications. Sometimes viscosity modifier, pour point depressant, addition of additives and chemical modification were utilized to meet the standard biolubricant studies were reported for general applications and non-edible oil

30

26, 27

24, 25

automotive

. Some reviews on

28

, drilling fluids

. However, the review of palm oil based biolubricant is still scanty. To the

best of our knowledge, current palm oil review focus on production of nanocarbon oil

32, 33

, biogas

34

29

and green fuel

35-37

31

, biocrude

. This article highlights state of art of catalyst for

isomerization of fatty acid. To further emphasize on specific feedstock, this study will elaborate on isomerization and opportunity of palm oil as biolubricant.

2. Isomerization of fatty acids to fuels and chemicals Fatty acid is the main substance in vegetable oil. This fatty acid require transformation to increase its functional application. The outcome of the industrialization made fatty acid as major feedstock for detergent, lubricant and surfactant. Fatty acid is a carboxylic acid with 12-20 carbon atoms. The properties of fatty acid, whether its composed of saturated and unsaturated, depend on its triglyceride source.

4 ACS Paragon Plus Environment

Page 5 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

One approach considers enhancing functionality of fatty acid is to convert linear of fatty acid to branched hydrocarbon over catalyst surface via isomerization. Isomerization could increase cetane number of fatty acids. In addition, this technique also successful to reduce cloud point of biodiesel production 38, 39, to convert coconut oil to jet fuel over Mo-Ni/ߛ-Al2O3 12 and to conduct etherification for biodiesel additive

40

Therefore, it is crucial to identify the promising

catalysts and to optimize the reaction condition for isomerization of fatty acid. Acidic supports such as zeolites and silicoaluminophosphates were among the first catalysts for isomerization. Silicoaluminophosphate (SAPO) has unique porous geometry with combination of medium acid proton. These properties favored hydroisomerization rather than cracking which initiate with β-scission. Incorporation of Pd to SAPO showed high conversion of sunflower, however this catalyst was easily deactivated 41. In another studies, Pd was also doped to natural bentonite which was activated with mineral acid at high temperature. The modified bentonite has high surface area and mesoporous. The hierarchical bentonites exhibited largest selectivity to isomerization due to presence of strong Brønsted acid site. Pd could be dispersed more homogeneously in the hierarchical bentonites with 28% dispersion. However, this catalyst more selective to linear paraffins 42. The isomerization studies were also conducted to upgrade soybean oil over Pt/SAPO-11. The SAPO-11 support has mesoporous aggregates at 3-10 nm, total surface area of 187 m2/g and 85% of Pt dispersion. The optimum condition was at 30 atm and 375-385 °C with isomerization selectivity up to 50%. This reaction occurred via decarbonylation and decarboxylation with little amount of aromatic and naphtenes 43. In another research, Pt was doped to SAPO-11 to convert with methyl palmitate at constant percentage namely at 3 wt.%

44

. The catalysts were prepared

by co-incipient wetness impregnation with various ratios of Sn to Pt and different temperatures.

5 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 45

It was shown that monometallic Pt was more active to C-C breakage. This mechanism was confirmed from large conversion of the feed to aliphatic carbon chain. The high conversion could be caused by the presence of dissociated hydrogen in Pt surface. Meanwhile, bimetallic PtSn-SAPO-11 in the same condition produced 4 times iso-palmitate than over monometallic Pt. This high isomer yield was attributed to presence of partially reduced tin oxides on platinum surface which reactive for hydrodeoxygenation. However, the amount of tin in the catalyst needs to optimize to prevent blockade of Pt active sites. The mechanistic reaction of Pt and PtSn over palmitate acid was illustrated in Fig. 1. In another research, it was revealed that Pt-SAPO-11 catalyst has low thermal stability for hydrotreating of vegetable oil. Therefore, the performance of this catalyst decreased. One potential method is to mix this catalyst with amine based surfactant. This technique could prevent desilication of its framework which maintain its acidity. Therefore, the catalyst performance became more stable

45

. In another research, several porous

materials such as DNL-G (RHO), SAPO-34 and hydrotalcite were studied to produce of nheptadecane as a product from decarboxylation of oleic acid. This study confirmed that SAPO34 exhibited the highest performance due to its highest acid site with opening ring of 0.38 nm. The isomerization reaction prefer relatively large pore diameter to accommodate fatty acids molecule 46.

6 ACS Paragon Plus Environment

Page 7 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

Fig. 1. Mechanistic reaction of Palmitic Acid over Pt and PtSn catalyst (adapted from Ref.45).

Stearic acid (C18:0) isomerization was evaluated over NiMo/γ-Al2O3-β-zeolites catalysts 47

. The catalysts were prepared with physically mixing zeolite and Al2O3. Therefore, certain

percentage of Ni and Mo salt were dissolved in citric acid. The reaction was carried out at 320 °C and 5 MPa in presence of tetralin. This study revealed that NiO and MoO3 tend to react with alkene pathways. Meanwhile, zeolite showed high FFA (free fatty acid) conversion. However, these catalysts still need further development due to their low isomerization yield.

7 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Zeolite based catalysts were also evaluated for fatty acid isomerization

Page 8 of 45

48

. The zeolites

were modified with Mo through witness impregnation method from their parent zeolites namely K-ZSM-22 and H-ZSM-22. The catalytic conversions of palmitic acid (C16:0) were performed at 4 MPa and 260 °C. The presence of H+ in H-ZSM-22 which smaller to K+ in K-ZSM-22 tend to ease penetration to zeolite acid site. Hence, the conversion of these parent zeolites was relatively higher over H-ZSM-22. This experiment also employed MoO3 and MoO2 to acquire better understanding to basic mechanism. This experiment proved that Mo4+ favored to HDO (hydrodeoxygenation), while Mo6+ favored to HDC (hydro-decarbonylation). Hence, it is become crucial concern to increase Mo4+/Mo6+ ratio to increase isomerization conversion and yield. Another promising catalysts for isomerization was platinum promoted tungstate modified zirconia

49

. These catalysts were divided to three types which depend on their preparation

techniques. This research verified that best prepared catalyst was zirconia from precipitation technique, which was followed by wet impregnation with tungsten salt. After calcination, platinum salt was added and recalcined at 500 °C. The sample of FFA in this experiment was palmitic acid (C16:0). The ratio of tungsten to Pt and operational condition for isomerization were also established in this experiment. The optimum conversion and selectivity were 79.1% and 89.9% respectively. Selected catalysts, which were studied for isomerization, are presented in Table 1.

8 ACS Paragon Plus Environment

Page 9 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47

Energy & Fuels

Table 1. Isomerization of fatty acid to fuel and chemicals. Catalyst Pt/SAPO-11 (Pt = 0.3wt%)

Operating condition T = 300oC; P = 3 MPa

Conversion 24.2%

Isomer yield 0.1wt%

Pt/SAPO-11 (Pt = 0.3wt%)

T = 375oC; P = 3 MPa

81.3%

12.1wt%

Pt-Sn/SAPO-11 (Pt = 0.3wt%; Sn = 0.3wt%)

T = 300oC; P = 3 MPa

22.9%

0.5wt%

Pt-Sn/SAPO-11 (Pt = 0.3wt%; Sn = 0.3wt%)

T = 375oC; P = 3 MPa

72.7%

31.4wt%

Pt-Sn/SAPO-11 (Pt = 0.3wt%; Sn = 0.6wt%)

T = 375oC; P = 3 MPa

86.6%

50.0wt%

NiMo/γ-Al2O3-βzeolite (β-zeolite = 10wt%; NiO + MoO3 =

T = 320oC; P = 5MPa

96%

Isomerization ratio = 0.5%

Selectivity/composition Selectivity: i-C15,16=0.3% n-C15,16=37.8% C16+ = 1.2% Palmitic acid = 27.7% Palmitylpalmitate = 20.2% Selectivity: i-C15,16=14.8% n-C15,16=50.4% C16+ = 1.5% Palmitic acid = 4.8% Palmitylpalmitate = 0% Selectivity: i-C15,16=2.0% n-C15,16=8.2% C16+ = 2.3% Palmitic acid = 22.5% Palmitylpalmitate = 48.6% Selectivity: i-C15,16=43.2% n-C15,16=16.1% C16+ = 2.9% Palmitic acid = 7.8% Palmitylpalmitate = 0.4% Selectivity: i-C15,16=57.8% n-C15,16=21.3% C16+ = 2.6% Palmitic acid = 3.6% Palmityl palmitate = 0.2% Selectivity: C16 = 5% C17 = 27% C19 = 69%

Remark Hydrogenation of methyl palmitate; WHSV = 5 h-1; H2/feedstock = 800 (v/v)

Ref 44

Hydrogenation of methyl palmitate; WHSV = 5 h-1; H2/feedstock = 800 (v/v)

44

The hydrogenation of methyl palmitate; WHSV = 5 h-1; H2/feedstock = 800 (v/v)

44

Hydrogenation of methyl palmitate; WHSV = 5 h-1; H2/feedstock = 800 (v/v)

44

Hydrogenation of methyl palmitate; WHSV = 5 h-1; H2/feedstock = 800 (v/v)

44

Hydrogenation of stearic acid; batch for 3 h; feed : stearic acid and tetralin

47

9 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47

20wt%; Mo + Ni= 20wt%) NiMo/γ-Al2O3-βzeolite (β-zeolite = 100wt%; NiO + MoO3 = 20wt%; Mo + Ni= 20wt%) K-ZSM-22

T = 320oC; P = 5MPa

62%

T = 260oC; P = 4MPa

63%

H-ZSM-22

T = 260oC; P = 4 MPa

95%

Mo/K-ZSM-22 (Mo : K-ZSM-22 = 0.5 (g:g))

T = 260oC; P = 4 MPa

100%

Mo/H-ZSM-22 (Mo : H-ZSM-22 = 0.5 (g:g))

T = 260oC; P = 4 MPa

100%

Pt/WO3/ZrO2 (W= 6.5 wt.%; Pt= 0.5 wt.% )

T= 220 °C, P= 160 psig of H2

79.1%

Isomerization ratio = 71%

71.1%

Page 10 of 45

Selectivity: C16 = 42% C17 = 44% C19 = 12%

Hydrogenation of stearic acid; batch for 3 h; feed : stearic acid and tetralin

47

Composition: C15 -COOH = 37.5% C15 -CHO = 8.2% n-C16= 17.5% iso-C16= 27.1% n-C15 = 9.7% Composition: C15 -COOH = 6.3% n-C16 = 17.8% iso-C16 = 12.0% n-C15 = 33.9% iso-C15 = 28.0% n-C14= 2.0% Composition: C15 -CHO = 25.1% n-C16 = 47.5% n-C15 = 27.4% Composition: n-C16 = 13.0% n-C15 = 34.9% iso-C15 = 47.9% n-C14 = 4.2%

Hydrogenation of palmitic acid; batch for 4 h; feed : palmitic acid and H2

48

Hydrogenation of palmitic acid; batch for 4 h; feed : palmitic acid and H2

48

Hydrogenation of palmitic acid; batch for 4 h; feed : palmitic acid and H2

48

Hydrogenation of palmitic acid; batch for 4 h; feed : palmitic acid and H2

48

i-C16= 89.9 %

WHSV = 1 h-1 Ratio mol of H2/n-C16= 2

49

10 ACS Paragon Plus Environment

Page 11 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

3. Isomerization of fatty acid with zeolites Zeolites are among potential catalysts for fatty acid isomerization. Zeolites have acid site inside their porosity which reactive to unsaturated carbon to form cyclic carbon chain. This cyclic chain contains unstable carbocations, tends to lose proton and release small molecules. This process sometimes followed with double bond migration to carbocyclic acid, which is preferable at high temperature reaction 50. Several zeolite types were studied for the conversion of palm oil 51. The zeolites consist of H-ZSM-5, H-β-Zeolite and H-USY. The research was intended to study of the effect of porosity for palm oil transformation to fuel. This experiment showed that H-ZSM-5 was the most superior catalyst to maximize gasoline. The porosity in H-ZSM-5 was the smallest than H-βZeolite and H-USY. Additionally, main factor to maximize the palm oil conversion in H-ZSM-5 was WHSV (weight hour space velocity) and optimum at 1 WHSV h-1. Meanwhile, increasing temperature and decreasing WHSV were key role to maximize palm oil conversion with βZeolite. In case of USY, the conversion increased drastically in the range of 400-450 °C. From this phenomenon, it was concluded that the large external surface area of zeolites was not the main factor for palm oil conversion instead of aluminum framework 51. Isomerization of oleic acid was investigated over beta zeolite. Different calcination temperatures were evaluated in the experiment in the range of 350-550 °C. The optimum temperature for calcination was at 400 °C with conversion of 67%. At above this temperature, the conversion decreased drastically due to the absence of Brønsted acid site. In addition, βZeolite should has mesoporous porosity which crucial for skeletal isomerization

52

. This

experiment was also identified effect of additives such as water, methanol and light hydrocarbons. This study confirmed that water behaved as good additives and water to oleic acid 11 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 45

ratio is 0.31. However, above this ratio, water could destruct zeolite structure and hard to recover. Meanwhile, methanol could ease ester formation which improve isomerization performance. However, this reaction requires addition of sulfuric acid to form an ester compound. The isomerization of ester pathways could be increased 5% by as compared with the absence of methanol. The last, addition of light hydrocarbon did not show crucial effect for oleic conversion 53. Ferrierite zeolites also exhibited high conversion of butene to iso-butene. But, this type of zeolites is challenging to recover

54

. Ferrierite has a distinctive characteristic to prevent coke

formation and undesired side reaction. The research to improve ferrierite performance was elaborated recently. These catalysts demonstrated high performance until 20 regeneration times without significant conversion and selectivity. The experiment was conducted to induce branch formation in stearic acid and followed with hydrogenation. H+-Ferrierite in the presence of distilled water showed conversion up to 80% at 260 °C for 4 h. The catalysts were showed excellent isomerization after heat regeneration until 4th times. After 5th and 6th usage, the catalysts required an acid-TPP (triphenylphosphine) treatment to retain their conversion and selectivity up to 19 times. This experiment also confirmed that substitution of NH4+ to ferrierite did not contribute significantly and still require same regeneration method 54. This technology is efficient to save cost until 51% as its still stable after 15 times 55. The isomerization of fatty acids with zeolites are listed in Table 2.

12 ACS Paragon Plus Environment

Page 13 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

Table 2. Isomerization of fatty acid with zeolites. Catalyst

Feed

Operating condition T= 350 °C

Conversion

HZM-5

Palm Oil

HZSM-5

Palm Oil

T= 400 °C

96.9%

Beta Zeolites

Palm Oil

T= 350 °C

82.2%

Beta zeolite

Palm Oil

T= 400 °C

86%

USY

Palm Oil

T= 350 °C

53.2%

USY

Palm Oil

T= 350 °C

61.5%

Beta zeolite

C18:1

250 °C, 0.3 MPa

Beta zeolite

C18:1

Ferrierite Zeolites with 1 mL water Ferrierite zeolite with 1 mL ethanol H+ Ferrierite zeolite regenerated with acid H+ Ferrierite zeolite regenerated with heat

Selectivity/composition

Remark

Ref

WHSV = 1 h-1

51

WHSV = 1 h-1

51

WHSV = 1 h-1

51

WHSV = 1 h-1

51

WHSV = 1 h-1

51

WHSV = 1 h-1

51

50%

Gasoline: 28.3% Kerosene: 9.1% Diesel: 5% Gasoline: 20.5% Kerosene: 11.4% Diesel: 11.4% Gasoline: 22% Kerosene: 9.4% Diesel: 9.2 Gasoline: 22.1 Kerosene: 12.1 Diesel: 10.7 Gasoline: 7.3 Kerosene: 9 Diesel: 10.6 Gasoline: 13 Kerosene: 13.8 Diesel: 0.8 Iso-C18:1= 58%

53

250 °C, 0.3 MPa

94%

Iso-C18:1= 80%

C18:1

250 °C, 40 Psi

94%

Methyl iso-C18: 85% Metyl C18: 6.2% C36: 2.3%

Zeolite loading= 2 g/100 g feed Zeolite loading = 5 g/100 g feed Zeolite loading: 2.5 wt.%

C18:1

250 °C, 40 Psi

95%

Metyl iso-C18: 85 % Metyl C18: 7.2% C36: 2.0%

Zeolite loading: 2.5 wt.%

54

C18:1

250 °C, 40 Psi

99%

Si/Al ratio: 17.5

54

C18:1

250 °C, 40 Psi

98%

Branch metyl esters: 84% Methyl C18: 7.1% γ-branch stearo lactone: 7.2% γ-stearo lactone:0.2% C36 metyl ester dimer: 1.5% Branch metyl esters: 81% Methyl stearate: 7.7% γ-branch stearo lactone: 7.5% γ-stearo lactone:0.3% C36 metyl ester dimer:

Si/Al ratio: 17.5

54

99%

53

54

13 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 45

3.5%

The hydrogenation of soybean oil was explored using ZSM-5 catalyst

56

. The catalysts

were prepared with competitive ion exchange (CIE) or incipient wet impregnation (IWI). This research revealed that CIE was better than IWI for dispersion of Pt with average particle size of 2.2 nm. In addition, Pt was located in small pore of ZSM-5 with pressure-jump IR analysis. The existence of Pt in near mouth of ZSM-5 was prospective to direct shape selective conversion of feed to specific product. The resulted ZSM-5 consist of mesopore and micropore volume after alkali treatment with NaOH. This catalyst was promising to enhance both low temperature fluidibility and oxidative stability for biolubricant production 56. Several zeolites such as ZSM-5, Beta Zeolite and Ferrierite were evaluated for isomerization of oleic acid recently

57

. The zeolites were initiated with heat treatment which

followed with proton exchange. The zeolites showed high performance in absence of TPP (triphenylphosphine) and low water content. Water is essential to prevent poisonous effect to acid site in carbonyl group and also generate new Brønsted acid site. The reaction was carried out at 260 °C for 8 h. This research confirmed that ferrierite as the best catalyst with 98% conversion and 80% selectivity. In addition, the catalyst also exhibited relatively stable performance in large scale installation with 96% conversion and 76% selectivity

57

. Another

research was focused to characterize acid site and accessibility of ferrierite for oleic acid isomerization 58. The five ferrierites were originated from commercial ferrierite with four similar Si/Al ratio. This research confirmed that there was no clear influence of Si/Al ratio for isomerization selectivity. The identification with NH3-TPD (Temperature Programmed Desorption) confirmed that ferrierite catalysts contained large density of Brønsted acid with very low Lewis acid site. The acid site was located near pore mouth in smaller channel, which could not be accessed with TPP (triphenylphosphine) according to observation with XPS and UV-Vis DRs. The highest selectivity for isomerization was achieved over ferrierite with low density of strong Brønsted acid site near 10 MR chain and low external acidity. This research also reported that rapid deactivation occurred due to pore mouth blockage and acid site deactivation 59. The isomerization of palmitic acid was evaluated over bifunctional Ni and Co zeolites. Isomerization could not occur in absence of metal over zeolite surface

60

. The composites were

prepared with solid state impregnation. Since the metal size was larger than zeolites channel, the metal was loaded on the surface of zeolites and only small amount of metal could enter to zeolite 14 ACS Paragon Plus Environment

Page 15 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

framework during calcination. The post-synthesis treatment was required to introduce mesoporous agent such TEAOH to obtain higher selectivity in isomerization. The best selectivity for isomerization was achieved at 260 °C and 4 MPa of H2 over Co based composites. High dispersion of Co could lead to better interaction with Co0 species and relatively strong acid site. Moreover, Co/zeolites demonstrated high stability after several usage

60

. Therefore, Co could

become promising substitution to rare-earth based metals.

4. Hierarchical Zeolites Hierarchical zeolites are unique types of zeolites with at least two different porosities. The porosity could be combination of micropores, mesopores or macropores. However, these two porous types must be interconnected with porous difference at least at one order of magnitude

61

. This characteristic was essential to enhance catalytic efficiency and overcome

diffusion constraint. Many types hierarchical zeolites have been reported such as MOR, MFI, BEA and MTW, which were classified according to ring number and framework. Hierarchical zeolites offer potential application such as alkylation over MTW (ZSM-12) 62, catalytic cracking of n-hexane over TON (ZSM-22) olefins with TON/MTT

65

63

and n-dodecane over BEA

, selective amidation with solvent free

64 66

, conversion of naphtha to and cracking light naphtha

with ZSM-48 67 Isomerization of fatty acid is a crucial step in biolubricant fabrication. The prospective catalyst for isomerization was based on ferrierite zeolites (FER). FER consists of perpendicular 10 MR (Membered Ring) and 8 MR. This types of zeolites have five Brønsted acid sites, which distributed in channel intersection, channel framework and hydroxyl group in 10 MR

68

.

However, it is still challenging to control this acid site in FER with desired characteristic. The first article to synthesis ferrierite was reported in the literature 69. Alkali treatment was utilized to prepare mesoporous ferrierite from commercial ferrierite (Si/Al = 29) without alteration effect on acidity and crystallinity of parent ferrierite. The addition of mesoporous site with desilication requires more harsh condition than other hierarchical zeolites. The final porosity of ferrierite was achieved until 3 to 4 time increment. Although the resulted ferrierites showed high performance for isomerization of 1-butene, but the chemicals still could not diffuse to 8 ring chain. The increasing of diffusion properties was reported with three post steps reaction from similar parent zeolites. The sequential steps required different reagents, namely NaAlO2, HCl and NaOH. The 15 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 45

addition of these chemicals was carried out at 80 °C. In the first step, the silica framework leaching occurred with the increase on mesoporous surface area and total volume with mesoporous pore around 10 nm. This characteristic was also followed by presence of amorphous alumina over its surface and decreased its acidity due to ionic exchange between H+ and Na+. In second step, amorphous alumina was removed with acid leaching, which also restored proton concentration and microporosity volume. This step increased amount of Brønsted acid and produced mesoporous site center at around 5 nm. In the final step, selective removal of Si debris was carried out with NaOH. The end product of this reaction has intercrystalline mesoporosity with better diffusion transfer and catalytic performance 70. The mesoporous generation of FER was also reported via subsequent dealumination and desilication process

71

. Aluminum leaching was performed with acid solution then followed by

alkali treatment. In the first step, Si/Al ratio was increased and produced high crystallinity zeolites. In this report, Al was not only presence in tetrahedral frameworks, but also in extraframework. Therefore, mesoporous creation was achieved in the next stage after alkali treatment, which was confirmed from increasing intensity of IR spectrum. This mesoporous generation induced easier access to active sites and efficient to design intrinsic acidity within microporous sites. The preparation of hierarchical micro and mesoporous material with better control of pore construction with appropriate acid strength is still a challenging topic in this decade. The synthesis procedures were based on delamination of layered zeolites 74

impregnation over mesoporous zeolite , recrystallization

75

72

, hard template

73

,

76

and partial dissolution . FER was

also successfully prepared with hydrothermal recrystallization in alkaline media. The alkaline solution was used to introduce mesoporosity to parent FER. At first, the FER framework would be partially dissolved in the medium and formed mesoporous site in the presence of a template such as CTMABr (cetyltrimethylammonium bromide). The concentration of alkaline solution was varied to obtain optimum selectivity for isomerization. Among the samples, the highest acid sites with an increase up to 76% than parent FER was achieved with the highest degree of crystallization. However, the highest isomerization selectivity over n-butene was accomplished with an intermediate crystallization. This was an indication that selectivity over FER could be increased with crystallization due to its medium acid strength and easier accessibility to acid site 77

. In another research, more uniform Pt distribution on FER catalyst was conducted with ion 16 ACS Paragon Plus Environment

Page 17 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

exchange. The platinum salt was tetraamineplatinum(II) nitrate under reflux. The better distribution of metal support could increase catalyst external surface area and enhance the hydrocracking performance 78. The preparation of FER was also investigated using organosilane as surfactant and 3(trimethoxysilyl) propyl octadecyl dimethyl ammonium chloride (TPOAC) as organic structure directing agent (OSDA)

79

. The synthesis method was based on hydrothermal with one step

approach. The resulted FER has mesoporous site with diameter of 6.5 nm which consist of intercrystalline and multilayer stacking. This research revealed that TPOAC content could specify the morphology and phase transition of FER. However, the crystallization of as synthesized FER was longer than the parent FER with a difference up to 12 h. In this study, all Al atoms were located in zeolite framework with lower acidity than the reference FER. The total acidity difference was at 44%, while the strong acid site decreased of 33%. The decrease of acidity would decrease isomerization, cracking and oligomerization

80

. These catalysts were

suitable for reacting bulky molecules of both reactant and product such as benzylation of toluene. In another research, FER was also produced with a combination of piperidine and tetramethylammonium hydroxide (TMAOH). The FER product has 3 times higher mesoporosity than reference FER. The concentration of TMAOH could differentiate morphology of FER with primary aggregates were in stacked nanosheets, while piperidinde was responsible for formation of FER framework. According to mass spectrometry, both of TMAOH and piperidine were exist in FER framework which correspond to ease generation of hierarchical zeolites. Moreover, with this method, Al atoms could be dispersed external surface area to enhance the catalytic properties 81

. Recently, FER synthesis was accomplished via simple hydrothermal using single OSDA

namely pyrrolidine. This method was highly influenced by crystallization temperature to promote nucleation and growth. This reaction favored low temperature namely at 120 °C. The FER has mesoporous structure and could be controlled with high alkalinity synthesis. FER was potential for butane isomerization due to its high selectivity and recovery 82. The preparation of FER was described in Table 3. The deactivation mechanism of FER was studied recently to enhance catalyst stability and regeneration. The sample was on oleic acid with shape selective catalysis. To identify active site for isomerization, TPP (triphenylphosphine) was chosen as probe as it could not enter FER 17 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 45

framework due to its larger size. TPP could neutralize acid site in FER external surface area and easier to localize responsible acid site for isomerization 83. In the presence of TPP, FER showed high selectivity for isomerization. However, in absence of TPP, oligomerization route favored as existence of larger external surface area. Meanwhile, the microporosity reduced significantly and decreased surface area in short period. Therefore, it was proposed that the high selectivity for isomerization was originated from the availability of the entrance of pores or pore mouth active sites. The capability of pore mouth of FER as active sites was also reported in the literature

84

.

Meanwhile, the deactivation of FER was attributed to polyunsaturated carbocation. This carbocation is important while the spent catalysts were washed with acetone or HCl. The acetone washing could decrease conversion significantly, while HCl could recover the activity of the catalyst to initial conversion. It was believed that the charged organic compound could be exist in FER, which exchangeable with acidic proton. In the other hand, the intermediate of unsaturated carbon consist of dodecyl benzene, which lead to pore blockage, and ketone which tend poison catalyst active site 85. The deactivation mechanism is illustrated in Figure 2.

18 ACS Paragon Plus Environment

Page 19 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

Figure 2. The mechanism of FER deactivation (adapted from Ref 85).

19 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47

Page 20 of 45

Table 3. Different hierarchical FER zeolites reported in open literatures. Materials Commercial H-ferrierite

OSDA -

Commercial ferrierite

-

T= 80 °C tNaAlO2 = 3 h tHCl = 3 h tNaOH = 0.5 h [HNO3] = 0.3 M [LiOH] = 0.2 M CTAMBr at 110 °C, 24 h [NaOH] : 0.41.8 M Pt content : 0.5 wt.% T= 250 °C TPOAC, T : 550 °C pyrrolidine t : 24 h

Intracrystalline Mesopore: 99 m2/g

Remark Increase total mesopore area to 3-4 times increment Increase metal dispersion to FER

Smeso : 108 m2/g Brønsted acid : 320 ߤ mol pyridine/g

Formation of mesoporous site in 8 MR

71

Brønsted acid : 99 ߤ mol pyridine/g

Medium acid strength for high selectivity to isomerization Increase external surface area

77

79

Na2SiO3

TMAOH, piperidine

T= 550 C t=6h [NH4+] for ion exchange = 1 M

Al2(SO4)3

pyrrolidine

T= 120 °C

Aggregates: 10-15 ߤm Thickness: < 50 nm Medium acid strength = 0.60 mmol NH3/g Particle size: 40-60

HF= 0.11 Suitable for bulky molecules HF = 0.052

WHSV = 16 h-1

82

Commercial ferrierite

Commercial ferrierite

Commercial ferrierite Na2SiO3

Condition T= 60-90 °C NaOH = 0.1-1 M

Characteristic Thin platelets: 50150 nm

V.meso: 0.143 cm3/g Brønsted acid : 76 ߤ mol pyridine/g Particle size: 8-16 ߤm

Ref 69

70

78

81

20 ACS Paragon Plus Environment

Page 21 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47

Energy & Fuels

Na2SiO3

nm Total pore volume: 0.461 cm3/g Notes: OSDA: organic structure-directing agent TPOAC: 3-(trimethoxysilyl)-propyl-octadecyl-dimethyl-ammoniumchloride TMAOH: tetramethylammonium hydroxide HF: Hierarchy factor; S: Surface; V: Volume

21 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 22 of 45

5. Optimizing reaction of fatty acid to bio-lubricants Lubricant is an essential substance especially for automotive application. Lubricant decreased the coefficient of surface friction which further prevent corrosion, reduce oxidation and sealing agent from unwanted contaminants

86

. Earlier lubricants made from mineral oil

contain zinc, magnesium, calcium and iron traces which induce faster corrosion

87

. Although

palm oil mill effluent (POME) could become safe additive to conventional lubricant, its performance at high temperature still doubted

88

. In addition, disposal of mineral oil to

environment could be harmful to marine and land ecosystem

89

. Therefore, the search of

alternative lubricant source rose as a blossom research in this few decades. The promising lubricant source is vegetable oil which has long chain fatty acids. This biolubricant has relatively better lubricity

90

, save to environment and biodegradable

91

. The

comparison of mineral and bio-based lubricant was studied comprehensively. Biolubricant displayed higher viscosity index than the mineral lubricant which potential to hydraulic fluids and gear oils

92

. The biolubricant was also promising for future eco-vehicle

87

. There are many

plants, which are potential as green lubricants. The selection of vegetable-based stocks could be different depend on availability in each country. Jatropha is one of the most prospective non-edible oils in Philippines, India, Indonesia, Thailand and Pakistan. Jatropha contain dominantly of oleic (C18:1) and linoleic acid (C18:2) and also potential for biodiesel

94, 95

93

. Since this type of vegetable oil comprise of unsaturated

fatty oil, the biolubricants production were initiated with epoxidation or saponification. Epoxidation technique is promising to produce biolubricants for high temperature applications. In this research, Jatropha oil was reacted with formic acid and hydrogen peroxide in assistance of sulfuric acid catalyst. The resulting epoxy jatropha oil was mixed with sodium metoxide for esterification. This study showed that epoxy jatropha has better lubricity than pure Jatropha oil. Finally, this experiment also demonstrated that diphenyl amine as one the most potential additives to jatropha based lubricants

96

. In another research, jatropha oil was mixed with

potassium hydroxide due to its high FFA content which lead to foam formation. Therefore, the methyl ester product was reacted with sodium metoxide to synthesize lubricants. This lubricants is potential for light gear applications

97

. Meanwhile, direct biolubricant production from

jatropha oil was also reported. Jatropha oil (JO) was mixed with trimethylolpropane (TMP) at 150 °C with molar ratio of JO and TMP was 4:1 in assistance of 2 wt.% of sulfuric acid catalyst. 22 ACS Paragon Plus Environment

Page 23 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

This reaction could convert 55% of JO and still need further improvement

98

. The effect of

temperature for conversion of Jatropha methyl ester (JME) to biolubricant was also studied recently. JME was reacted with TMP over sodium metoxide as catalyst. This reaction occurred via second order reaction rate with an activation energy 3.94 kJ/mol. To maintain forward reaction, the ratio of JMP to TMP was 3.9:1 and optimum condition for conversion was at 150 °C and 10 mbar 99. Preparation of bio-lubricants from Castor oil was also investigated recently. Castor oil comprises dominantly of ricinoleic acid (C18:1) which also consist of hydroxyl group. The presence of hydroxyl group tends to increase boiling point and viscosity due to formation of hydrogen bond. The production of castor based bio-lubricant was reported by commercial resins as catalyst. The three different resins namely Dowex 50W-X8, Amberlyst-15, Purolite CT275DR and reacted with various types and ratios of alcohols. These resins were based on styrene-divynyl benzene with different on loading percentages. Amberlyst-15 showed the highest performance for biolubricant conversions. This result was correlated to the highest Brønsted acid site in Amberlyst-15 due to the rich of –SO3H group. The total of acid amount in Amberlyst-15 was 6.03 meq H+/g. The optimum condition for biolubricant conversion was at 100 °C for 4 h with an assistance of 2-ethyl-hexanol as solvent

100

. Another biolubricant based on castor oil was also

reported recently. The catalyst was also based on cation metal exchange resin with high Brønsted acid. In addition, the type of resin was modified to be macroreticular to enhance thermal resistance and sulphur leaching. This method could convert 95% of Castor oil in 4 h. Finally, the resulted biolubricant was also better than the feed basestocks oil 101. Soybean consist dominantly of linoleic acid (C18:2) and oleic acid (C18:1)

102

. Epoxidized

soybean was one of potential candidate for biolubricant at high temperature application

103

. The

presence of acid catalyst could open carbon ring group which could be followed with easy functionalization with nucleophilic compound. The most common catalyst for biolubricant conversion were sulphuric acid and acetic acid. However, these catalysts were difficult to recover and corrosive to reactor. Therefore, the development of heterogeneous acid catalyst become a crucial concern to the development of biolubricant. The advancement to acid catalysts were reported recently using SAC-13 which contain fluorosulfonic acid polymer. After reaction with catalyst, oxirane intermediate was reacted with nucleofil for functionalization. This research utilized methanol, ethanol, 2-propanol and 2-butanol as sample of nucleofil. This study showed 23 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 24 of 45

that 2-butanol was the best performance for oxirane conversion. Finally, the optimum condition was at 4 wt.% of catalyst and 80 °C. The mechanism of carbon ring opening is presented in Figure 3

104

. The epoxidized soybean showed higher viscosity than conventional lubricants.

Finally, the overall working efficiency of pure epoxy soybean based lubricant exhibited similar properties to mineral lubricants 105.

H

R1

O

H R2

H

R

H O

H

R2

OH R2

SO3H

Florosulf onic acid

R1

R1

H

O

H

R-OH

R Open Alcohol Carbon ring Nucleof il Figure 3. Mechanistic for carbon opening of with an acid catalyst (adapted from Ref 104).

Epoxy Reactants

The current challenges for biolubricant from vegetable oils were related to its properties such as poor low temperature properties and narrow viscosity. Therefore, it is important to identify factor which hinder previous obstacles. Some methods such as vary length of alcohol, number and position of double band, molecular polarity and branching. However, the correlation between chemical structure and physical characteristic was still scanty. The design of the desired biolubricant properties was explored recently by varying spacer between two ester of moieties. The C18-n-C18 diester were used as sample with various n lengths. From this research, it was confirmed that crystallization performance affected to steric repulsion while n < 4. For even number at n>4, the viscosity was linear with the increase of n. Meanwhile, for odd number of n, the viscosity was at large range due to the presence of specific molecular arrangement 106. Oleic acid (C18:1) is the largest fatty acid constituent in palm oil. Direct conversion of oleic acid to biolubricant was reported by Araujo et al

107

at 350 °C and 1 atm using H3PO4 and

H3PO4/Al2O3 as catalysts for biofuel and biolubricant production. H3PO4 was impregnated to Al2O3. The overall porosity of Al2O3 decreased with the incorporation of phosphoric acid. It was confirmed that H3PO4 located in inner surface of Al2O3. This study showed that Al2O3 has best

24 ACS Paragon Plus Environment

Page 25 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

selectivity to biolubricant (C19-C22) due to its highest Brønsted and Lewis acid sites. The catalyst selectivity was up to 95%. The conversion of palm oil to biolubricant was also reported recently over base catalysts. The catalysts were based on CaO with different loadings of SrO. This research confirmed that the increase of SrO content could increase metal oxides basicity. Increasing basicity was not favored for lubricant formation due to competition with side reaction namely saponification. Therefore, the highest conversion was achieved at the lowest SrO content. This research also established optimum condition for biolubricant preparation namely at 180 °C and 2 mbar. The proposed mechanistic reaction for biolubricant formation was as follows: (1) alcohol and ester adsorb at catalyst surface which contain Ca, Sr and O as active sites. (2) The neighboring alcohol and ester at catalyst surface were occur as intermediate condition. (3) These intermediate states would react with ester and produce biolubricant. These overall reaction would require precise palm oil to TMP ratio and occur continuously until –OH group was eliminated

108

. The

mechanistic reaction of biolubricant formation was presented in Figure 4. The biolubricant formation from vegetables oil is listed in Table 4.

25 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 26 of 45

Figure 4. The mechanism of palm oil based bio-lubricant with base catalyst (adapted form Ref. 108

26 ACS Paragon Plus Environment

Page 27 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47

Energy & Fuels

Table 4. Biolubricant formation from vegetables oil. Source Castor Oil

Catalyst Condition Dowex 50W-X8 T= 100 °C

Conversion 1 h: 13% 4 h: 23% 1 h: 90% 4 h: 100% 1 h: 25% 4 h: 80% Not reported

Castor Oil

Amberlyst-15

T= 100 °C

Castor Oil

T= 100 °C

Jatropha Oil

Purolite CT275DR Sodium metoxide

T= 120 °C; t= 2.5 h

Jatropha Oil

Sodium metoxide

T= 60-65 °C

Not reported

Epoxidize Jatropha

Sodium metoxide

T= 60-65 °C

Not reported

Jatropha Oil

T= 150 °C; t= 3 h

55%

Palm Oil

TrimetylolPropane (TMP) CaO5SrO

90%

Palm Oil

CaO10SrO

Palm Oil

CaO15SrO

Palm Oil

CaO20SrO

POME

CH3ONa

T= 180 °C; P= 2 mbar T= 180 °C; P= 2 mbar T= 180 °C; P= 2 mbar T= 180 °C; P= 2 mbar T = 140 °C; P = 25 mbar

Properties PP: - 39 °C V (100 °C): 7.77 V.index : 132

Remark Catalyst loading: 10 wt% 2-ethyl- hexanol: castor oil = 2:1

Ref 100

100

100

PP: -7 °C V (100 °C): 10.96 V.index: 195 PP: - 6 °C V (100 °C): 7.9 V.index: 205 PP: 0 °C V(100 °C): 18.2 V.index:139 PP: -30 °C V : 79 cP FP: > 300 °C Not reported

Catalyst loading: 0.8 % w/w Jatropha oil:metha-nol = 3.5:1

99

Catalyst loading: 1 wt.%

96

Catalyst loading: 1 wt.%

97

Catalyst loading: 2 % w/w Jatropha: TMP= 4:1

98

Catalyst loading: 1% w/w

108

85%

Not reported

Catalyst loading: 1% w/w

108

83%

Not reported

Catalyst loading: 1% w/w

108

81%

Not reported

Catalyst loading: 1% w/w

108

Catalyst Loading: 1% v/v POME: TMP = 4:1

109

94.6%

V (100 °C)= 9 cSt V.index = 176 PP = -2 °C

27 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47

Page 28 of 45

Notes: PP: Pour point; FP: Flash Point; V: Viscosity; V.index : Viscosity index; POME: Palm oil methyl esters.

28 ACS Paragon Plus Environment

Page 29 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

The transesterification of POME (Palm Oil Methyl Ester) and TMP (Trimetylolpropane) for lubricant formation was occurred reversibly with three steps of reaction. These paths were monitored with gas chromatography (GC) which placed during the reaction. The study was carried out at 110 °C and 1.2 mbar. The observation showed that the first (mono-ester) and second reaction (diester) occurred only 1 min. However, the overall reaction (tri-ester) need until 1 h to achieve the equilibrium state. Therefore, the kinetics for the reaction were created for future optimization to show rate constant and energy activation

110

. The same approach to

produce biolubricant based palm oil was also elaborated using Oscillatory Flow Reactor (OFR). This method could convert up to 94% of POME for 25 min. This method is faster than conventional stirred reactor due to better global mixing which also increase radical velocity of reactant. However, the biolubricant properties still not meet ISO standard 109. POME consist dominantly of methyl oleate namely at 75%. POME was also reported as basestocks for biolubricant using calcium metoxide as catalysts. Calcium metoxide more selective to transesterification rather than saponification reaction. The temperature was varied at 140-190 °C. High temperature facilitated faster transesterification reaction, but too high temperature could lead to decompose of POME. This research revealed that, 170 °C was optimum temperature for biolubricant formation. The conversion reached up to 98% at TMP: POME ratio 1:6, pressure 50 mbar and catalyst loading 0.3 wt.% 111. Zeolites could become potential additive for palm oil based lubricant to replace harmfull additive in biolubricant industry. The zeolites which prepared with free organic templates namely Na-FAU, Na-EMT and K-LTL. The average crystal size of Na-FAU, Na-EMT, K-LTL were 15 nm, 20 nm, 30 nm respectively. This sample of zeolites were mixed with biolubricant and monitored for 10, 20, 30 and 40 days. The zeolites disperse completely on biolubricant with clear solution. From the observation, Na-FAU showed best performance to prevent formation of solid black polymer, peroxides and polimerization of palm oil. These conclusions were derived from FT-IR and 1H NMR comparison of blank biolubricant and addition of 0.5 wt.% of zeolites. The propose mechanistic zeolites for additive were due to selective absorption of peroxide compound to hydrophilic framework in Na-FAU. In addition, the intermediate of oxidized product was also stabilized due to strong interaction between C=C and O-O-H group. Finally, the three dimensional framework of Na-FAU zeolites was also contributed to ease palm oil penetration 112. 29 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 30 of 45

Zeolites were also showed excellent performance for palmitic acid esterification. Palmitic acid is second main component in palm oil

113

. The zeolites consist of commercial zeolites

namely ZSM-5 and H-Y which were calcined at 600 °C for 24 h prior to use. The best zeolite catalyst was H-Y-60 with conversion 100% for 3 h at ratio of methanol to palmitic acid of 2:1. This result was twofold faster than previous reported catalyst, TPA/MCM-41 with much lower of methanol ratio. This research confirmed that no clear correlation between acid number to esterification reaction. The main factors of esterification were balancing hydrophilic/ hydrophobicity and porosity. The hydrophobicity has linier correlation with Si/Al ratio. The higher Si/Al ratio, the higher the hydrophobicity. Therefore, the water product tend to desorb on the surface and increase forward esterification reaction. Meanwhile, the microporosity on H-Y60 was more selective to methanol due to their unique partial charge distributions and dipole moment. This catalyst also showed promising recyclability after five usage without significant activity reduction 114. The production of biolubricants was also achieved with coupling fatty acid. This method could transform directly of carboxyclic acid to paraffin with two consecutive reactions namely ketonic decarboxylation and hydrogenation. Ketone was produced with base catalyst namely MgO with high conversion and selectivity at 97% and 95%, respectively. Therefore, the optimum hydrogenation was accomplished with Pt/MgO with selectivity up to 70%. These two processes were separated to two bed reactors to prevent reductive decarboxylation of acid to the metal component. In addition, the article also reported alternative support to Pt namely Al2O3 with lower temperature than MgO and yield C10-C23 of 90% 115. The alternative method on biolubricant synthesis was reported recently with reductive etherification. The industrial biolubricant production was employed cationic oligomerisation with corrosive catalyst such as AlCl3, HF and BF3. The recent method react alcohol with either aldyhyde or ketone in presence of 2.5 mol% of Pd/C catalyst. High molecular weight of alcohol was fabricated from hydrogenation of vegetable oil, while carbonyl groups were generated by fermentation of sugars. The biolubricant yield has low volatility and viscosity and decreased the oil refill and has similar characteristics with industrial biolubricants

116

. The mechanism of

current biolubricant production is presented in Figure 5.

30 ACS Paragon Plus Environment

Page 31 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

Figure 5. Reductive Etherification for Biolubricant Production (adapted from Ref.

116

).

31 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 32 of 45

6. Conclusions, challenges, and the ways forward Isomerization is an important path to improve physicochemical properties of biolubricant from fatty acids. Potential catalysts have unique number of mesoporosity site such as SAPO. The isomerization performance increased in presence of metal such Pt and bimetallic metal, PtSn. The metal should be dispersed uniformly on the porous substrate, which was achieved with coincipient wetness impregnation. In addition, the metal content must be optimized to prevent blockage of metal active sites. Zeolites could become alternative catalyst for isomerization. The existence of Brønsted acid associated with tunable mesoporous sites exhibited high selectivity towards isomerization. The β-zeolites could be modified with proton exchange and produce higher isomerization performance than USY and H-ZSM-5. However, among the zeolites, Ferrierites (FER) is the best candidate for isomerization. The FER showed the best performance attributed to low external acidity and strong Brønsted acid near 10 MR with active sites on FER pore mouth. Therefore, the FER showed excellent stability after washing with acid-TPP (triphenyl phosphine) which stable until 19th repetition. The direct synthesis of FER was achieved with one OSDA (organic structure directing agent) namely pyrrolidine. Meanwhile, the combination of acid and base leaching could be utilized to enhance metal dispersion on commercial FER. The metal dispersion was essential to increase external surface area. However, the FER external surface area was not responsible for high isomerization activity. FER selectivity for isomerization favored mesopore site near 10 MR and 8 MR. Some studies on biolubricant production from vegetable oil were also highlighted. The vegetable oil could be originated from soybean, castor, jatropha and palm oil. Meanwhile, the catalyst for biolubricant production could be varied from homogenous such as sulphuric acid or heterogeneous catalyst. Palm oil based biolubricants were reported via transesterification method using based catalyst namely calcium oxide. Base catalysis is faster than acid catalysis and also less corrosive to industrial apparatus. However, too basic catalyst was not favorable due to harsh competition with saponification as side reaction. Calcium oxide could be doped with strontium and convert 90% of palm oil to biolubricant with addition of 1 wt.% of catalyst. Future work could be directed to produce branch fatty acid with isomerization using hierarchical zeolites and

32 ACS Paragon Plus Environment

Page 33 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

doped with metal as basic site. In addition, new industrial method such as reductive etherification could be optimized and propose cheaper catalysts than Pd/C. The forecast of increasing demand of biolubricant must be followed with rapid development of biolubricant formulation. Existing biolubricant still require further improvement to increase oxidative stability and poor low temperature characteristic. This biolubricants tend to form solid residue at low temperature which further decrease lubricant movement

117

. In

addition, the biolubricant, which contain unsaturated carbon was easy to oxidize become ketone and carboxylic acid wear

119

118

. Meanwhile, the presence of free fatty acids could lead to corrosion and

. Therefore, it was essential to develop certain easy, safe and low-cost method for

production of biolubricant. Finally, the guidelines of biolubricant testing, tribological and physicochemical were vital role for the large-scale production of biolubricants.

Acknowledgements

A.M. would like to express appreciation for the support of MJIIT student incentive. The authors would like to acknowledge the funding provided by King Abdulaziz City for Science and Technology (KACST) through the Science & Technology Unit in the Center of Research Excellence in Nanotechnology at King Fahd University of Petroleum and Minerals (KFUPM) for supporting this work through project No. 13-NAN1702-04 as part of the National Science, Technology and Innovation Plan.

33 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 34 of 45

References 1. Sheil, D.; Casson, A.; Meijaard, E.; van Noordwjik, M.; Gaskell, J.; Sunderland-Groves, J.; Wertz, K.; Kanninen, M., The impacts and opportunities of oil palm in Southeast Asia: What do we know and what do we need to know? 2009. 2. USAID, Biofuels in Asia: An Analysis of Sustainability Option. In USAID, Ed. Bangkok, 2009. 3. WEC, Biofuels: Policies, Standards and Technologies. In World Energy Council, London, Ed. Council WE: 2010. 4. USDA In Indonesian Oilseeds and Products: Annual Report 2008., Jakarta, 2012; Report, G., Ed. Jakarta, 2012. 5. Mukherjee, I.; Sovacool, B. K., Palm oil-based biofuels and sustainability in southeast Asia: A review of Indonesia, Malaysia, and Thailand. Renewable and Sustainable Energy Reviews 2014, 37, 1-12. 6. Gan, P. Y.; Li, Z. D., Econometric study on Malaysia ‫׳‬s palm oil position in the world market to 2035. Renewable and Sustainable Energy Reviews 2014, 39, 740-747. 7. Hansen, S., Feasibility Study of Performing an Life Cycle Assessment on Crude Palm Oil Production in Malaysia (9 pp). The International Journal of Life Cycle Assessment 2007, 12, (1), 50-58. 8. Ahmad, S.; Kadir, M. Z. A. A.; Shafie, S., Current perspective of the renewable energy development in Malaysia. Renewable and Sustainable Energy Reviews 2011, 15, (2), 897-904. 9. Wicke, B.; Sikkema, R.; Dornburg, V.; Faaij, A., Exploring land use changes and the role of palm oil production in Indonesia and Malaysia. Land Use Policy 2011, 28, (1), 193-206. 10. Shomchoam, B.; Yoosuk, B., Eco-friendly lubricant by partial hydrogenation of palm oil over Pd/γ-Al2O3 catalyst. Industrial Crops and Products 2014, 62, 395-399. 11. Obidzinski, K.; Andriani, R.; Komarudin, H.; Andrianto, A., Environmental and social impacts of oil palm plantations and their implications for biofuel production in Indonesia. Ecology and Society 2012, 17, (1), 25. 12. Kimura, T.; Imai, H.; Li, X.; Sakashita, K.; Asaoka, S.; Al-Khattaf, S. S., Hydroconversion of Triglycerides to Hydrocarbons Over Mo–Ni/γ-Al2O3 Catalyst Under Low Hydrogen Pressure. Catalysis Letters 2013, 143, (11), 1175-1181. 13. Van Aelst, J.; Philippaerts, A.; Turner, S.; Van Tendeloo, G.; Jacobs, P.; Sels, B., Heterogeneous conjugation of vegetable oil with alkaline treated highly dispersed Ru/USY catalysts. Applied Catalysis A: General 2016, 526, 172-182. 14. Zhang, S.; Zhang, Z.; Steichen, D. Skeletal isomerization of fatty acids. 6831184, 2004/12//, 2004. 15. Tomifuji, T.; Abe, H.; Matsumura, Y.; Sakuma, Y. Process for the preparation of branched chain fatty acids and alkyl esters thereof. 5677473, 1997/10//, 1997. 16. Kenneally, C. J.; Connor, D. S. Process for the branching of saturated and/or unsaturated fatty acids and/or alkyl esters thereof. 6455716, 2002/09//, 2002. 17. Shuguang, Z.; Zongchao, Z. Fatty acid isomerization with mesoporous zeolites. 6723862, 2004/04//, 2004. 18. Kasza, T.; Kalló, D.; Hancsók, J., Quality improvement of bio-paraffin mixtures. Fuel 2014, 120, 17. 19. Echaroj, S.; Santikunaporn, M.; Chavadej, S., Transformation of Bioderived 1-Decanol to Diesellike Fuel and Biobased Oil via Dehydration and Oligomerization Reactions. Energy & Fuels 2017, 31, (9), 9465-9476. 20. Yang, Y.; Wang, Q.; Zhang, X.; Wang, L.; Li, G., Hydrotreating of C18 fatty acids to hydrocarbons on sulphided NiW/SiO2–Al2O3. Fuel Processing Technology 2013, 116, 165-174. 21. Ameen, M.; Azizan, M. T.; Yusup, S.; Ramli, A.; Yasir, M., Catalytic hydrodeoxygenation of triglycerides: An approach to clean diesel fuel production. Renewable and Sustainable Energy Reviews 2017, 80, 1072-1088.

34 ACS Paragon Plus Environment

Page 35 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

22. Hancsók, J.; Eller, Z.; Pölczmann, G.; Varga, Z.; Holló, A.; Varga, G., Sustainable production of bioparaffins in a crude oil refinery. Clean Technologies and Environmental Policy 2014, 16, (7), 14451454. 23. Maher, K. D.; Bressler, D. C., Pyrolysis of triglyceride materials for the production of renewable fuels and chemicals. Bioresource Technology 2007, 98, (12), 2351-2368. 24. McNutt, J.; He, Q., Development of biolubricants from vegetable oils via chemical modification. Journal of Industrial and Engineering Chemistry 2016, 36, 1-12. 25. Rizvi, S. Q. A., A comprehensive review of lubricant chemistry, technology, selection, and design. ASTM International: West Conshohocken, PA, 2009; p 657. 26. Syahir, A. Z.; Zulkifli, N. W. M.; Masjuki, H. H.; Kalam, M. A.; Alabdulkarem, A.; Gulzar, M.; Khuong, L. S.; Harith, M. H., A review on bio-based lubricants and their applications. Journal of Cleaner Production 2017, 168, 997-1016. 27. Salimon, J.; Abdullah, B. M.; Yusop, R. M.; Salih, N., Synthesis, reactivity and application studies for different biolubricants. Chemistry Central Journal 2014, 8, (1), 16. 28. Mobarak, H. M.; Niza Mohamad, E.; Masjuki, H. H.; Kalam, M. A.; Al Mahmud, K. A. H.; Habibullah, M.; Ashraful, A. M., The prospects of biolubricants as alternatives in automotive applications. Renewable and Sustainable Energy Reviews 2014, 33, 34-43. 29. Kania, D.; Yunus, R.; Omar, R.; Abdul Rashid, S.; Mohamad Jan, B., A review of biolubricants in drilling fluids: Recent research, performance, and applications. Journal of Petroleum Science and Engineering 2015, 135, 177-184. 30. Singh, Y.; Farooq, A.; Raza, A.; Mahmood, M. A.; Jain, S., Sustainability of a non-edible vegetable oil based bio-lubricant for automotive applications: A review. Process Safety and Environmental Protection 2017, 111, 701-713. 31. Misson, M.; Haron, R.; Kamaroddin, M. F. A.; Amin, N. A. S., Pretreatment of empty palm fruit bunch for production of chemicals via catalytic pyrolysis. Bioresource Technology 2009, 100, (11), 28672873. 32. Hosseini, S. E.; Wahid, M. A., Utilization of palm solid residue as a source of renewable and sustainable energy in Malaysia. Renewable and Sustainable Energy Reviews 2014, 40, 621-632. 33. Sumathi, S.; Chai, S. P.; Mohamed, A. R., Utilization of oil palm as a source of renewable energy in Malaysia. Renewable and Sustainable Energy Reviews 2008, 12, (9), 2404-2421. 34. Loh, S. K.; Nasrin, A. B.; Mohamad Azri, S.; Nurul Adela, B.; Muzzammil, N.; Daryl Jay, T.; Stasha Eleanor, R. A.; Lim, W. S.; Choo, Y. M.; Kaltschmitt, M., First Report on Malaysia’s experiences and development in biogas capture and utilization from palm oil mill effluent under the Economic Transformation Programme: Current and future perspectives. Renewable and Sustainable Energy Reviews 2017, 74, 1257-1274. 35. Mekhilef, S.; Siga, S.; Saidur, R., A review on palm oil biodiesel as a source of renewable fuel. Renewable and Sustainable Energy Reviews 2011, 15, (4), 1937-1949. 36. Hameed, B. H.; Lai, L. F.; Chin, L. H., Production of biodiesel from palm oil (Elaeis guineensis) using heterogeneous catalyst: An optimized process. Fuel Processing Technology 2009, 90, (4), 606-610. 37. Lim, S.; Teong, L. K., Recent trends, opportunities and challenges of biodiesel in Malaysia: An overview. Renewable and Sustainable Energy Reviews 2010, 14, (3), 938-954. 38. Reaume, S.; Ellis, N., Use of Isomerization and Hydroisomerization Reactions to Improve the Cold Flow Properties of Vegetable Oil Based Biodiesel. Energies 2013, 6, (12), 619-633. 39. Reaume, S. J.; Ellis, N., Use of hydroisomerization to reduce the cloud point of saturated fatty acids and methyl esters used in biodiesel production. Biomass and Bioenergy 2013, 49, 188-196. 40. Yang, F.; Zhang, S.; Zhang, Z. C.; Mao, J.; Li, S.; Yin, J.; Zhou, J., A biodiesel additive: etherification of 5-hydroxymethylfurfural with isobutene to tert-butoxymethylfurfural. Catalysis Science & Technology 2015, 5, (9), 4602-4612. 35 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 36 of 45

41. Kikhtyanin, O. V.; Rubanov, A. E.; Ayupov, A. B.; Echevsky, G. V., Hydroconversion of sunflower oil on Pd/SAPO-31 catalyst. Fuel 2010, 89, (10), 3085-3092. 42. Detoni, C.; Bertella, F.; Souza, M. M. V. M.; Pergher, S. B. C.; Aranda, D. A. G., Palladium supported on clays to catalytic deoxygenation of soybean fatty acids. Applied Clay Science 2014, 95, 388395. 43. Herskowitz, M.; Landau, M. V.; Reizner, Y.; Berger, D., A commercially-viable, one-step process for production of green diesel from soybean oil on Pt/SAPO-11. Fuel 2013, 111, 157-164. 44. Chen, N.; Ren, Y.; Qian, E. W., Elucidation of the active phase in PtSn/SAPO-11 for hydrodeoxygenation of methyl palmitate. Journal of Catalysis 2016, 334, 79-88. 45. Rabaev, M.; Landau, M. V.; Vidruk-Nehemya, R.; Goldbourt, A.; Herskowitz, M., Improvement of hydrothermal stability of Pt/SAPO-11 catalyst in hydrodeoxygenation–isomerization–aromatization of vegetable oil. Journal of Catalysis 2015, 332, 164-176. 46. Ahmadi, M.; Nambo, A.; Jasinski, J. B.; Ratnasamy, P.; Carreon, M. A., Decarboxylation of oleic acid over Pt catalysts supported on small-pore zeolites and hydrotalcite. Catalysis Science & Technology 2015, 5, (1), 380-388. 47. Huang, Z.; Ding, S.; Li, Z.; Lin, H.; Li, F.; Li, L.; Zhong, Z.; Gao, C.; Chen, C.; Li, Y., Catalytic conversion of stearic acid to fuel oil in a hydrogen donor. International Journal of Hydrogen Energy 2016, 41, (37), 16402-16414. 48. Shi, Y.; Cao, Y.; Duan, Y.; Chen, H.; Chen, Y.; Yang, M.; Wu, Y., Upgrading of palmitic acid to isoalkanes over bi-functional Mo/ZSM-22 catalysts. Green Chemistry 2016, 18, (17), 4633-4648. 49. Zhang, S.; Zhang, Y.; Tierney, J. W.; Wender, I., Hydroisomerization of normal hexadecane with platinum-promoted tungstate-modified zirconia catalysts. Applied Catalysis A: General 2000, 193, (1), 155-171. 50. Huang, X.; Zhu, J.; Lin, Z., β-hydrogen elimination of five-membered-ring metallacycles. Is it possible? Organometallics 2004, 23, (17), 4154-4159. 51. Twaiq, F. A.; Zabidi, N. A. M.; Bhatia, S., Catalytic Conversion of Palm Oil to Hydrocarbons: Performance of Various Zeolite Catalysts. Industrial & Engineering Chemistry Research 1999, 38, (9), 3230-3237. 52. Zhang, S.; Zhang, Z. C., Skeletal isomerization of unsaturated fatty acids: the role of mesopores in HBeta zeolites. Catalysis Letters 2007, 115, (3-4), 114-121. 53. Ha, L.; Mao, J.; Zhou, J.; Zhang, Z. C.; Zhang, S., Skeletal isomerization of unsaturated fatty acids on Beta zeolites: Effects of calcination temperature and additives. Applied Catalysis A: General 2009, 356, (1), 52-56. 54. Ngo, H. L., Improved zeolite regeneration processes for preparing saturated branched-chain fatty acids: Zeolite synthesis of saturated branched-chain fatty acids. European Journal of Lipid Science and Technology 2014, 116, (5), 645-652. 55. Ngo, H. L.; Yee, W. C.; McAloon, A. J.; Haas, M. J., Techno-Economic Analysis of an Improved Process for Producing Saturated Branched-Chain Fatty Acids. Journal of Agricultural Science 2014, 6, (10). 56. Van Aelst, J.; Philippaerts, A.; Bartholomeeusen, E.; Fayad, E.; Thibault-Starzyk, F.; Lu, J.; Schryvers, D.; Ooms, R.; Verboekend, D.; Jacobs, P.; Sels, B., Towards biolubricant compatible vegetable oils by pore mouth hydrogenation with shape-selective Pt/ZSM-5 catalysts. Catalysis Science & Technology 2016, 6, (8), 2820-2828. 57. Sarker, M. I.; Latona, R. J.; Moreau, R. A.; Micheroni, D.; Jones, K. C.; Ngo, H., Convenient and Environmentally Friendly Production of Isostearic Acid with Protonic Forms of Ammonium Cationic Zeolites. European Journal of Lipid Science and Technology 2017, 119, (12), 1700262. 58. L., N. H.; Alberto, N.; Wenbin, L.; A., F. T., Zeolite-catalyzed isomerization of oleic acid to branched-chain isomers. European Journal of Lipid Science and Technology 2007, 109, (3), 214-224. 36 ACS Paragon Plus Environment

Page 37 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

59. Wiedemann, S. C. C.; Muñoz-Murillo, A.; Oord, R.; van Bergen-Brenkman, T.; Wels, B.; Bruijnincx, P. C. A.; Weckhuysen, B. M., Skeletal isomerisation of oleic acid over ferrierite: Influence of acid site number, accessibility and strength on activity and selectivity. Journal of Catalysis 2015, 329, 195-205. 60. Shi, Y.; Xing, E.; Cao, Y.; Liu, M.; Wu, K.; Yang, M.; Wu, Y., Tailoring product distribution during upgrading of palmitic acid over bi-functional metal/zeolite catalysts. Chemical Engineering Science 2017, 166, 262-273. 61. Hartmann, M.; Machoke, A. G.; Schwieger, W., Catalytic test reactions for the evaluation of hierarchical zeolites. Chemical Society Reviews 2016, 45, (12), 3313-3330. 62. Sanhoob, M. A.; Muraza, O.; Tago, T.; Taniguchi, T.; Watanabe, G.; Masuda, T., Development of mesoporous ZSM-12 zeolite and its application in alkylation of 2-methylnaphthalene. Research on Chemical Intermediates 2016, 42, (7), 6437-6448. 63. Jamil, A. K.; Muraza, O.; Ahmed, M. H.; Zainalabdeen, A.; Muramoto, K.; Nakasaka, Y.; Yamani, Z. H.; Yoshikawa, T.; Masuda, T., Hydrothermally stable acid-modified ZSM-22 zeolite for selective propylene production via steam-assisted catalytic cracking of n-hexane. Microporous and Mesoporous Materials 2018, 260, 30-39. 64. Ahmed, M. H. M.; Muraza, O.; Jamil, A. K.; Shafei, E. N.; Yamani, Z. H.; Choi, K.-H., Steam Catalytic Cracking of n-Dodecane over Ni and Ni/Co Bimetallic Catalyst Supported on Hierarchical BEA Zeolite. Energy & Fuels 2017, 31, (5), 5482-5490. 65. Lateef, S. A.; Bakare, I. A.; Muraza, O., Microwave assisted synthesis of MTT-TON intergrowth crystals for the catalytic conversion of naphtha to olefins. Microporous and Mesoporous Materials 2018, 260, 253-259. 66. Mäki-Arvela, P.; Zhu, J.; Kumar, N.; Eränen, K.; Aho, A.; Linden, J.; Salonen, J.; Peurla, M.; Mazur, A.; Matveev, V.; Murzin, D. Y., Solvent-free “green” amidation of stearic acid for synthesis of biologically active alkylamides over iron supported heterogeneous catalysts. Applied Catalysis A: General 2017, 542, 350-358. 67. Ahmed, M. H. M.; Muraza, O.; Nakaoka, S.; Jamil, A. K.; Mayoral, A.; Sebastian, V.; Yamani, Z. H.; Masuda, T., Stability Assessment of Regenerated Hierarchical ZSM-48 Zeolite Designed by Post-Synthesis Treatment for Catalytic Cracking of Light Naphtha. Energy & Fuels 2017, 31, (12), 14097-14103. 68. Zholobenko, V. L.; Lukyanov, D. B.; Dwyer, J.; Smith, W. J., Ferrierite and SUZ-4 Zeolite:  Characterization of Acid Sites. The Journal of Physical Chemistry B 1998, 102, (15), 2715-2721. 69. 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. 70. Verboekend, D.; Caicedo-Realpe, R.; Bonilla, A.; Santiago, M.; Pérez-Ramírez, J., Properties and Functions of Hierarchical Ferrierite Zeolites Obtained by Sequential Post-Synthesis Treatments. Chemistry of Materials 2010, 22, (16), 4679-4689. 71. Brylewska, K.; Tarach, K. A.; Mozgawa, W.; Olejniczak, Z.; Filek, U.; Góra-Marek, K., Modification of ferrierite through post-synthesis treatments. Acidic and catalytic properties. Journal of Molecular Structure 2016, 1126, 147-153. 72. Corma, A.; Fornes, V.; Pergher, S. B.; Maesen, T. L. M.; Buglass, J. G., Delaminated zeolite precursors as selective acidic catalysts. Nature 1998, 396, 353. 73. Jacobsen, C. J. H.; Madsen, C.; Houzvicka, J.; Schmidt, I.; Carlsson, A., Mesoporous Zeolite Single Crystals. Journal of the American Chemical Society 2000, 122, (29), 7116-7117. 74. Liu, Y.; Zhang, W.; Pinnavaia, T. J., Steam-Stable MSU-S Aluminosilicate Mesostructures Assembled from Zeolite ZSM-5 and Zeolite Beta Seeds. Angewandte Chemie International Edition 2001, 40, (7), 1255-1258. 75. On, D. T.; Kaliaguine, S., Large-pore mesoporous materials with semi-crystalline zeolitic frameworks. Angewandte Chemie - International Edition 2001, 40, (17), 3248-3251. 37 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 38 of 45

76. Inagaki, S.; Ogura, M.; Inami, T.; Sasaki, Y.; Kikuchi, E.; Matsukata, M., Synthesis of MCM-41-type mesoporous materials using filtrate of alkaline dissolution of ZSM-5 zeolite. Microporous and Mesoporous Materials 2004, 74, (1), 163-170. 77. Khitev, Y. P.; Kolyagin, Y. G.; Ivanova, I. I.; Ponomareva, O. A.; 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, (1-3), 201-207. 78. Fuentes-Ordóñez, E. G.; Salbidegoitia, J. A.; Ayastuy, J. L.; Gutiérrez-Ortiz, M. A.; GonzálezMarcos, M. P.; González-Velasco, J. R., High external surface Pt/zeolite catalysts for improving polystyrene hydrocracking. Catalysis Today 2014, 227, 163-170. 79. 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. 80. Lee, Y.; Park, M. B.; Kim, P. S.; Vicente, A.; Fernandez, C.; Nam, I.-S.; Hong, S. B., Synthesis and Catalytic Behavior of Ferrierite Zeolite Nanoneedles. ACS Catalysis 2013, 3, (4), 617-621. 81. 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. 82. Chu, W.; Li, X.; Zhu, X.; Xie, S.; Guo, C.; Liu, S.; Chen, F.; Xu, L., Size-controlled synthesis of hierarchical ferrierite zeolite and its catalytic application in 1-butene skeletal isomerization. Microporous and Mesoporous Materials 2017, 240, 189-196. 83. Wichterlová, B.; Žilkova, N.; Uvarova, E.; Čejka, J.; Sarv, P.; Paganini, C.; Lercher, J. A., Effect of Broensted and Lewis sites in ferrierites on skeletal isomerization of n-butenes. Applied Catalysis A: General 1999, 182, (2), 297-308. 84. Claude, M. C.; Martens, J. A., Monomethyl-Branching of Long n-Alkanes in the Range from Decane to Tetracosane on Pt/H-ZSM-22 Bifunctional Catalyst. Journal of Catalysis 2000, 190, (1), 39-48. 85. Wiedemann, S. C. C.; Stewart, J. A.; Soulimani, F.; van Bergen-Brenkman, T.; Langelaar, S.; Wels, B.; de Peinder, P.; Bruijnincx, P. C. A.; Weckhuysen, B. M., Skeletal isomerisation of oleic acid over ferrierite in the presence and absence of triphenylphosphine: Pore mouth catalysis and related deactivation mechanisms. Journal of Catalysis 2014, 316, 24-35. 86. Panchal, T. M.; Patel, A.; Chauhan, D. D.; Thomas, M.; Patel, J. V., A methodological review on bio-lubricants from vegetable oil based resources. Renewable and Sustainable Energy Reviews 2017, 70, 65-70. 87. Miller, A. L.; Stipe, C. B.; Habjan, M. C.; Ahlstrand, G. G., Role of Lubrication Oil in Particulate Emissions from a Hydrogen-Powered Internal Combustion Engine. Environmental Science & Technology 2007, 41, (19), 6828-6835. 88. Maleque, M. A.; Masjuki, H. H.; Haseeb, A. S. M. A., Effect of mechanical factors on tribological properties of palm oil methyl ester blended lubricant. Wear 2000, 239, (1), 117-125. 89. Ssempebwa, J. C.; Carpenter, D. O., The generation, use and disposal of waste crankcase oil in developing countries: A case for Kampala district, Uganda. Journal of Hazardous Materials 2009, 161, (23), 835-841. 90. Salimon, J.; Salih, N.; Yousif, E., Chemically modified biolubricant basestocks from epoxidized oleic acid: Improved low temperature properties and oxidative stability. Journal of Saudi Chemical Society 2011, 15, (3), 195-201. 91. Bartz, W. J., Lubricants and the environment. Tribology international 1998, 31, (1), 35-47. 92. Regueira, T.; Lugo, L.; Fandiño, O.; López, E. R.; Fernández, J., Compressibilities and viscosities of reference and vegetable oils for their use as hydraulic fluids and lubricants. Green Chemistry 2011, 13, (5), 1293. 38 ACS Paragon Plus Environment

Page 39 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

93. Emil, A.; Yaakob, Z.; Satheesh Kumar, M. N.; Jahim, J. M.; Salimon, J., Comparative Evaluation of Physicochemical Properties of Jatropha Seed Oil from Malaysia, Indonesia and Thailand. Journal of the American Oil Chemists' Society 2010, 87, (6), 689-695. 94. Mofijur, M.; Masjuki, H. H.; Kalam, M. A.; Hazrat, M. A.; Liaquat, A. M.; Shahabuddin, M.; Varman, M., Prospects of biodiesel from Jatropha in Malaysia. Renewable and Sustainable Energy Reviews 2012, 16, (7), 5007-5020. 95. Wang, R.; Hanna, M. A.; Zhou, W.-W.; Bhadury, P. S.; Chen, Q.; Song, B.-A.; Yang, S., Production and selected fuel properties of biodiesel from promising non-edible oils: Euphorbia lathyris L., Sapium sebiferum L. and Jatropha curcas L. Bioresource Technology 2011, 102, (2), 1194-1199. 96. Sammaiah, A.; Padmaja, K. V.; Narayna Prasad, R. B., Synthesis of Epoxy Jatropha Oil and its Evaluation for Lubricant Properties. Journal of Oleo Science 2014, 63, (6), 637-643. 97. S, B., Production of biolubricant from Jatropha curcas seed oil. Journal of Chemical Engineering and Materials Science 2013, 4, (6), 72-79. 98. Arbain, N. H.; Salimon, J., Synthesis and characterization of ester trimethylolpropane based Jatropha curcas oil as biolubricant base stocks. Journal of Science and Technology 2010, 2, (2). 99. Gunam Resul, M. F. M.; Mohd. Ghazi, T. I.; Idris, A., Kinetic study of jatropha biolubricant from transesterification of jatropha curcas oil with trimethylolpropane: Effects of temperature. Industrial Crops and Products 2012, 38, 87-92. 100. Saboya, R. M. A.; Cecilia, J. A.; García-Sancho, C.; Sales, A. V.; de Luna, F. M. T.; RodríguezCastellón, E.; Cavalcante, C. L., Assessment of commercial resins in the biolubricants production from free fatty acids of castor oil. Catalysis Today 2017, 279, 274-285. 101. Saboya, R. M. A.; Cecilia, J. A.; García-Sancho, C.; Sales, A. V.; de Luna, F. M. T.; RodríguezCastellón, E.; Cavalcante, C. L., Synthesis of biolubricants by the esterification of free fatty acids from castor oil with branched alcohols using cationic exchange resins as catalysts. Industrial Crops and Products 2017, 104, 52-61. 102. Ramos, M. J.; Fernández, C. M.; Casas, A.; Rodríguez, L.; Pérez, Á., Influence of fatty acid composition of raw materials on biodiesel properties. Bioresource Technology 2009, 100, (1), 261-268. 103. Adhvaryu, A.; Erhan, S. Z., Epoxidized soybean oil as a potential source of high-temperature lubricants. Industrial Crops and Products 2002, 15, (3), 247-254. 104. Turco, R.; Tesser, R.; Vitiello, R.; Russo, V.; Andini, S.; Serio, M. D., Synthesis of Biolubricant Basestocks from Epoxidized Soybean Oil. Catalysts 2017, 7, (10), 309. 105. Ting, C.-C.; Chen, C.-C., Viscosity and working efficiency analysis of soybean oil based biolubricants. Measurement 2011, 44, (8), 1337-1341. 106. Raghunanan, L.; Narine, S. S., Engineering Green Lubricants I: Optimizing Thermal and Flow Properties of Linear Diesters Derived from Vegetable Oils. ACS Sustainable Chemistry & Engineering 2016, 4, (3), 686-692. 107. Araujo, L. R. R. d.; Scofield, C. F.; Pastura, N. M. R.; Gonzalez, W. d. A., H3PO4/Al2O3 catalysts: characterization and catalytic evaluation of oleic acid conversion to biofuels and biolubricant. Materials Research 2006, 9, 181-184. 108. Ivan-Tan, C. T.; Islam, A.; Yunus, R.; Taufiq-Yap, Y. H., Screening of solid base catalysts on palm oil based biolubricant synthesis. Journal of Cleaner Production 2017, 148, 441-451. 109. Koh, M. Y.; Mohd. Ghazi, T. I.; Idris, A., Synthesis of palm based biolubricant in an oscillatory flow reactor (OFR). Industrial Crops and Products 2014, 52, 567-574. 110. Hamid, H. A.; Yunus, R.; Rashid, U.; Choong, T. S. Y.; Al-Muhtaseb, A. a. H., Synthesis of palm oilbased trimethylolpropane ester as potential biolubricant: Chemical kinetics modeling. Chemical Engineering Journal 2012, 200-202, 532-540.

39 ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 40 of 45

111. Chang, T.-S.; Masood, H.; Yunus, R.; Rashid, U.; Choong, T. S. Y.; Biak, D. R. A., Activity of Calcium Methoxide Catalyst for Synthesis of High Oleic Palm Oil Based Trimethylolpropane Triesters as Lubricant Base Stock. Industrial & Engineering Chemistry Research 2012, 51, (15), 5438-5442. 112. Tan, K.-H.; Awala, H.; Mukti, R. R.; Wong, K.-L.; Ling, T. C.; Mintova, S.; Ng, E.-P., Zeolite nanoparticles as effective antioxidant additive for the preservation of palm oil-based lubricant. Journal of the Taiwan Institute of Chemical Engineers 2016, 58, 565-571. 113. Yunus, R.; Fakhru'l-Razi, A.; Ooi, T. L.; Iyuke, S.; Idris, A., Development of optimum synthesis method for transesterification of palm oil methyl esters and trimethylolpropane to environmentally acceptable palm oil-based lubricant. 2003; Vol. 15, p 35-41. 114. Prinsen, P.; Luque, R.; González-Arellano, C., Zeolite catalyzed palmitic acid esterification. Microporous and Mesoporous Materials 2018, 262, 133-139. 115. Corma, A.; Renz, M.; Schaverien, C., Coupling Fatty Acids by Ketonic Decarboxylation Using Solid Catalysts for the Direct Production of Diesel, Lubricants, and Chemicals. ChemSusChem 2008, 1, (8-9), 739-741. 116. Jadhav, D.; Grippo, A. M.; Shylesh, S.; Gokhale, A. A.; Redshaw, J.; Bell, A. T., Production of Biomass-Based Automotive Lubricants by Reductive Etherification. ChemSusChem 2017, 10, (11), 25272533. 117. Erhan, S. Z.; Asadauskas, S., Lubricant basestocks from vegetable oils. Industrial crops and products 2000, 11, (2), 277-282. 118. Shahbazi, K.; Mehta, S. A.; Moore, R. G.; Ursenbach, M. G., The Effect of Oxidation on Viscosity of Oil-Based Drilling Fluids. 2006. 119. Mannekote, J. K.; Kailas, S. V., The Effect of Oxidation on the Tribological Performance of Few Vegetable Oils. Journal of Materials Research and Technology 2012, 1, (2), 91-95.

40 ACS Paragon Plus Environment

Page 41 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41

Energy & Fuels

ACS Paragon Plus Environment Fig. 1. Mechanistic reaction of Palmitic Acid over Pt and PtSn catalyst (adapted from Ref. [44])

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41

ACS Paragon Plus Environment

Figure 2. The Mechanism of FER deactivation (adapted from Ref [83]).

Page 42 of 45

Page 43 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41

Energy & Fuels

Figure 3. Mechanistic for carbon opening of with an acid catalyst (adapted from Ref [102]).

ACS Paragon Plus Environment

Energy & Fuels 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41

Figure 4. The mechanism of palm oil based bio-lubricant with base catalyst (adapted form Ref. [106]). ACS Paragon Plus Environment

Page 44 of 45

Page 45 of 45 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41

Energy & Fuels

ACS Paragon Plus Environment

Figure 5. Reductive Etherification for Biolubricant Production (adapted from Ref. 115).