Role of Solvent in Catalytic Conversion of Oleic Acid to Aviation

(5) Therefore, the development of alternative liquid fuels in aviation is becoming increasingly important ...... 1987, 91 (2) 496– 500 DOI: 10.1021/...
0 downloads 0 Views 634KB Size
Subscriber access provided by Binghamton University | Libraries

Article

Role of Solvent in Catalytic Conversion of Oleic Acid to Aviation Biofuels Qiurong Tian, Zihao Zhang, Feng Zhou, Kequan Chen, Jie Fu, Xiuyang Lu, and Pingkai Ouyang Energy Fuels, Just Accepted Manuscript • Publication Date (Web): 22 May 2017 Downloaded from http://pubs.acs.org on May 24, 2017

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

Energy & Fuels 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

Role of Solvent in Catalytic Conversion of Oleic Acid to Aviation Biofuels Qiurong Tiana†, Zihao Zhanga†, Feng Zhoub, Kequan Chenc, Jie Fua*, Xiuyang Lua, Pingkai Ouyanga,c a

Key Laboratory of Biomass Chemical Engineering of Ministry of Education, College of

Chemical and Biological Engineering, Zhejiang University, Hangzhou 310027, China b

Fushun Research Institute of Petroleum and Petrochemicals, SINOPEC, Fushun 113001,

China c

State Key Laboratory of Materials-Oriented Chemical Engineering, College of

Biotechnology and Pharmaceutical, Nanjing Tech University, Nanjing 211816, China †

These authors contributed equally to this work.

* Corresponding author Jie Fu, Tel: +86 571 87951065 E-mail address: [email protected] Abstract: The role of solvents in the conversion of oleic acid over Pt/C was studied. Three solvent systems (solvent-free, water, and dodecane) were employed for the conversion of oleic acid over Pt/C at 350 °C. Decarboxylation, hydrogen transfer, and

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

aromatization were observed in these three reaction systems. Compared to the non-solvent reaction system, much slower decarboxylation and aromatization rates and fewer heptadecane and aromatic products were observed in the hydrothermal and dodecane reaction systems.

The decarboxylation and aromatization rates and yield of

heptadecane and aromatics decreased with increased dodecane loading in the dodecane reaction system, and the decarboxylation and aromatization rates and yield of heptadecane and aromatics significantly decreased with the increase of water in the hydrothermal reaction system.

The effects of solvent loading, catalyst loading, and reaction time on the

reactions (decarboxylation, hydrogen transfer, and aromatization) were investigated. The reaction behaviors of 1-heptadecene with different solvents were studied, and N2 adsorption-desorption and thermogravimetric analysis of fresh and spent Pt/C in the three reaction systems was also performed. The results indicate that the competition of dodecane for the Pt/C active sites is mainly responsible for the slow decarboxylation and aromatization rates.

In addition to the similar influencing factor to that in the dodecane

system, H+ released from water and hydrogen bonding, which inhibited the ionization of carboxyl groups, was the key influencing factor for the slower decarboxylation and aromatization rates obtained under hydrothermal conditions.

ACS Paragon Plus Environment

Page 2 of 45

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

Keywords: Oleic acid; Pt/C; Dodecane; Water; Non-solvent;

1. Introduction Renewable and green liquid fuel sources are being aggressively explored owing to depleting petroleum reserves, increased fuel demand and environmental concerns, and one of the vital fuel sources is biofuels 1.

Bioethanol and biodiesel are currently the

most prominent biofuels in commercial production and use, and biodiesel can be further de-oxygenated (upgraded) to be an environmental friendly alternative liquid fuel for aviation fuel 2-4. Jet fuel, with real prices almost tripling from approximately $1.30/gallon in 2000 to approximately $3.00/gallon in 2012, has accounted for approximately 10% of the U.S. petroleum refinery production over the past two decades 5. In addition to the challenge of the increasing and volatile price of jet fuel, the aviation industry faces environmental concerns associated with aviation fuel, including its impact on air quality and greenhouse gas emission 5.

Therefore, the development of alternative

liquid fuels in aviation is becoming increasingly important and useful 6. Aviation fuel consists of a complex mixture of C8-C17 paraffins, aromatics and naphthenes; as a result, many researchers are interested in the synthesis of

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 4 of 45

biomass-derived long-chain alkanes and aromatics, which can be further refined into jet fuel.

Fatty acids and their derivatives, which can be extracted from animal fat and

vegetable oil in large quantities, have been regarded as candidates for producing long-chain alkanes and aromatics, respectively 7-13. Lercher and co-workers 14 have reported a route to convert crude microalgae oil into diesel-range alkanes over heterogeneous catalysts. The selectivity for paraffins and aromatics in the conversion of triglycerides has also been comprehensively studied by changing the distribution of Lewis and Bronsted sites of hierarchical catalysts 15. However, most studies focus only on the production of long-chain paraffins from fatty acids 16-18. For unsaturated fatty acids, the conversions of unsaturated fatty acid in the hydrothermal, organic solvent and non-solvent system have been reported in the previous studies. For non-solvent system, in 2008, Murzin and co-workers

19

first reported the

successful deoxygenation of unsaturated fatty acids over Pt/C. In 2012, Na et al.20 found that oleic acid can be converted to heptadecane and heptadecene with a total yield of 72% over 5 wt% Pt/C in the solvent-free system, and aromatics were also detected. In 2015, Carron et al.21 have reported that the selectivity for heptadecane from oleic acid reached 66% under solvent-free condition using a Pt-SAPO-34 catalyst.

In 2016, Fu et al.22

systematically studied the reaction pathway and mechanism for the conversion of oleic acid to heptadecane and aromatics. For the system of dodecane, catalytic deoxygenation

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

of oleic acid in dodecane was performed by Lamb and co-workers 23, the results showed that only 12% heptadecane was obtained without added H2 over Pt/C from the conversion of oleic acid, but no data about the armatics was reported. Murzin et al.24 have also studied on the conversion of oleic acid in dodecane over 1 wt% Pd/C under the atmosphere of 1% hydrogen in argon. The products were stearic acid and little aromatics. Fu and co-workers

25

suggested that it was not clear whether hydrocarbon solvents (e.g.,

dodecane) play a key role in activity and selectivity for fatty acid decarboxylation. Thereafter, in 2011, Fu et al. proposed a new approach to convert saturated and unsaturated fatty acids to hydrocarbons in near or supercritical water, and a less than 20% selectivity for heptadecane from unsaturated fatty acids was obtained over Pt/C after 2.5 h without added H2. In 2015, Yeh et al.

26

discovered Pt3Sn/C showed higher selectivity

for the decarboxylation of oleic acid with the selectivity to heptadecane of 60%. The decarboxylation of fatty acids in organic solvent and water and without solvent has been previously reported, and numerous systematic studies27-30 on the influence of reaction time, temperature, pressure, atmosphere, catalyst and support on the decarboxylation of fatty acids have been performed over many years.

However, to the best of our

knowledge, the different reaction behaviors in dodecane, hydrothermal and non-solvent system on the conversion of oleic acid at the same reaction conditions (reactors, catalysts, reaction atmosphere etc.) were never compared before.

In addition, the influence factor

of different solvent on the conversion of oleic acid is unclear and not systematically studied. The aim of this work is to demonstrate the role of solvents (water, dodecane and

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

solvent-free) on the conversion of oleic acid using Pt/C. The effects of solvent loading, catalyst loading and reaction time along with the reaction of 1-heptadecene with different solvents were investigated. Additionally, the catalytic activity and stability was also examined to determine the relationship between the solvent and the catalyst. We observed different phenomena when using different solvents and, for the first time, deduced the role of solvent in the conversion of oleic acid.

2. Experimental section 2.1 Materials Oleic acid (>99.0% purity), undecylbenzene (>98% purity) and 1-heptadecene (>99.5% purity) were purchased from TCI. Acetone (analytic reagent grade) was obtained from Sinopharm Chemical Reagent Co., Ltd. Heptadecane and dodecane (>98% purity) were obtained from Aladdin Industrial Corporation. Commercial 5% Pt/C and stearic acid (>98.5% purity) were obtained from Sigma–Aldrich, USA. All of the above chemicals and the catalyst were used as received. 2.2 Experimental procedure The oleic acid conversion was carried out in a micro-batch reactor (1.67 cm3 volume)

ACS Paragon Plus Environment

Page 6 of 45

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

assembled from one 3/8-in. tube and two 3/8-in. caps, and the reactors were purchased from Swagelok, USA. In a typical experiment, 75 mg of reactant was added to the reactor, followed by 15 mg Pt/C and 0.75 mL solvent. After being loaded, each reactor was sealed by attaching and then tightening the reactor cap. Then, the reactor was placed in a fluidized sand bath (TECHNE SBL-2D) that was already at the desired reaction temperature. As soon as the reaction time reached the desired value, the reactor was submerged in ambient temperature water to quench the reaction. The sample in the cooled reactor was rinsed with acetone and transferred into a centrifuge tube, the reactor was rinsed until the total volume reached 15 mL, and after being filtered, the sample was injected into a gas chromatography (GC) vial for analysis. 2.3 Characterization N2 adsorption-desorption was performed at 77 K in a static volumetric apparatus (Micromeritics 3Flex).

Samples were degassed at 300 °C for 10 h before N2 adsorption.

The specific surface area was determined by the Brunauer-Emmett-Teller (BET) equation, and the pore size and volume were calculated according to the Barrett-Joyner-Halenda (BJH) method.

All calculations were achieved using the adsorption isotherms.

The thermogravimetric characteristics were measured using TGA (TA-Q500). The

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 8 of 45

fresh and spent catalyst were heated to 600 °C at 10 °C min-1 with flowing air at 50 cm3 min-1. In order to detect the depletion of Pt, the liquid sample after reaction was rinsed with acetone until the total volume reached 15 mL, and then 1 mL sample was transferred to a volumetric flask, diluted with water until the total volume was 100 mL.

The metal in

the solution was determined by the inductively coupled plasma mass spectrometry, X Series II (Thermo Fisher Scientific). 2.4 Analysis method The samples were analyzed by a gas chromatograph (GC, Agilent 7890B) equipped with a 30 m × 0.25 mm × 0.25 µm HP-5MS capillary column, a mass spectrometer (Agilent 5977A MSD), a flame ionization detector (FID) and a thermal conductivity detector (TCD).

A 1-µL sample was injected into the GC at a split ratio of 10:1, where

the temperature of the injector was 280 °C and the carrier gas (nitrogen) flow rate was 11.383 mL min-1.

The remaining sample was split into three equal portions, which

flowed to the MS, FID and TCD.

The temperature of these three detectors was 280 °C.

The FID temperature was 280 °C with an H2 flow of 30 mL min-1, an air flow of 400 mL min-1, and a makeup N2 flow of 25 mL min-1.

The MS had a solvent delay of 2.75 min

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 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

and scanned masses from 50.00 to 300 amu.

The oven temperature program consisted

of a 4-min soak at 50 °C, followed by a 20 °C min-1 ramp to 250 °C and was finally held for 2 min.

The reaction products were identified by fragmentation patterns from the MS

detector and by calibration with known standards.

Quantitative analysis of each

compound was completed using calibration curves obtained from FID signals. The reactant molar conversion was calculated as the moles of oleic acid consumed divided by the initial moles of oleic acid loaded into the reactor. Yields were calculated as the moles of product recovered divided by the theoretical moles of product while oleic acid is consumed completely.

Uncertainties reported herein are standard deviations,

which were determined by replicating the experiments.

Each data point represents the

mean result from three independent experiments.

3. Results and discussion 3.1 Reaction behavior of oleic acid with different solvents The reaction behavior in non-solvent, hydrothermal, dodecane reaction systems was studied. An oleic acid loading of 75 mg, Pt/C loading of 15 mg, reaction temperature of 350 °C and reaction time of 80 min were used in the experiments in Section 3.1. Table S1

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 10 of 45

showed the performance of 14 different kinds of catalysts on the decarboxylation and aromatization of oleic acid. Pt/C showed the best catalytic performance on the decarboxylation and aromatization of oleic acid, and then Pt/C was chosen as the model catalyst. Figure 1 shows a typical GC/FID chromatogram of the in-situ hydrogen transfer, decarboxylation/decarbonylation and aromatization products from oleic acid under the condition of non-solvent. The major products were heptadecane and aromatics, and the aromatics were mainly identified as 1-methydecylbenzene, 1,1-dimethylnonylbenzene, undecylbenzene and 2-undercylphenol.

In these systems, similar kinds of aromatics

were detected, but the yields of aromatics were significantly different.

Figure 2 shows

the conversion of oleic acid in non-solvent, hydrothermal and dodecane solvent systems over Pt/C. Figure S1 shows the conversion of oleic acid in methanol, formic acid and cyclohexane solvent systems over Pt/C. For the solvent-free reaction, a 71% yield of heptadecane and 19% yield of aromatics were obtained. When 0.75 mL of dodecane was added as the solvent, a 41% yield of heptadecane, 4% yield of aromatics and 31% yield of stearic acid were obtained. When 0.75 mL of water was added as the solvent, only a yield of 12% heptadecane, 3% aromatics and 68% stearic acid was obtained. The results indicate that decarboxylation, hydrogen transfer, and aromatization occurred in these three

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

reaction systems.

However, the yield of heptadecane and aromatics both decreased

remarkably with the addition of dodecane and water. The addition of solvent did not accelerate the decarboxylation and aromatization rates, but restricted the reactions of decarboxylation and aromatization. To determine the reason for the restrained conversion of oleic acid by the solvent, a series of experiments were carried out and described in the following sections.

3.2 Effect of solvent loading To determine the influence of the solvent, the conversion of oleic acid was performed with different solvent loadings. The same oleic acid loading (75 mg), Pt/C loading (15 mg), reaction temperature (350 °C) and reaction time (80 min) as before were used in the experiments in Section 3.2. The solvent loadings ranged from 0 to 1 mL. Figure 3 shows the effect of solvent loading on the conversion of oleic acid. Figure 3a shows that the effect of dodecane loading on the conversion of oleic acid was not significant. As the loading of dodecane increased from 0 mL to 1 mL, the yield of hydrogenated product (stearic acid) increased from 0% to 83%; the yield of heptadecane decreased from 71% to 13.3%; and the yield of aromatics decreased from 19% to 2%.

ACS Paragon Plus Environment

It indicates that the

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

in-situ hydrogen transfer was not influenced by the increase of dodecane, but decarboxylation and aromatization rates were restrained. The influence way of dodecane might be compete with oleic acid for the active sites of the catalyst. The more dodecane was added, the less active sites remained for oleic acid. Therefore, the decarboxylation and aromatization rates were slower, and the yields of aromatics and heptadecane were also lower, and the reaction was detained at the hydrogenated product (stearic acid). Figure 3b shows that the effect of water loading on the conversion of oleic acid was significant. As the loading of water increased from 0 mL to 0.25 mL, the yield of hydrogenated product (stearic acid) increased from 0% to 79%; the yield of heptadecane decreased from 71% to 18%; and the yield of aromatics decreased from 18% to 2%. In contrast, and the yields of heptadecane and aromatics kept stable when the loading of water increased from 0.25 mL to 1mL.

It indicates that decarboxylation and

aromatization rates were also restrained under the hydrothermal condition, similar with the reaction behavior in dodecane.

Furthermore, the inhibition effect of water was much

more significant than that of dodecane.

The remarkable influence of water on the

reaction of oleic acid can be attributed to the following possible reasons. One is the competition for the active sites of the catalyst, and another is that the existence of H+

ACS Paragon Plus Environment

Page 12 of 45

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

restricted the ionization of carboxyl groups, which may be the key impacting factor. Furthermore, Shurvell et al.31 reported that a hydrogen bonding exists between acetic acid and water.

Therefore, a hydrogen bonding may also exist between oleic acid and water,

and this hydrogen bonding may inhibit the ionization of carboxyl groups, further slowing the decarboxylation and aromatization rates in the hydrothermal system.

To further

prove that H+ released from water and hydrogen bonding existing between oleic acid and water is the influence factor, the effect of solvent on the reaction of 1-heptadecene without carboxyl group was investigated in Section 3.5. 3.3 Effect of catalyst loading To determine the influencing factors and causes, the conversion of oleic acid was conducted with different catalyst loadings. The same oleic acid loading (75 mg), solvent loading (0.75 mL), reaction temperature (350 °C) and reaction time (80 min) as before were used in the experiments Section 3.3. The Pt/C loading ranged from 5 mg to 20 mg. Figure 4 shows the effect of catalyst loading on the conversion of oleic acid with different solvents (a: non-solvent 22; b: dodecane; c: water). Under the condition of non-solvent, the loading of catalyst did not have significant influence on the yields of heptadecane and aromatics, which has been reported in our previous research 22 (Figure 4a).

ACS Paragon Plus Environment

When

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

dodecane was used as the solvent, the decarboxylation and aromatization rates and yields of heptadecane and aromatics obviously increased with increased Pt/C loading, especially that the yield of heptadecane increased from 10.1% to 58.1% (Figure 4b). It proves that the supposition that dodecane competed for active sites is reasonable. With the increase of catalyst loading, the competition from dodecane for the active sites of the catalyst decreased. When water was used as the solvent, the decarboxylation and aromatization rates and yield of aromatics and heptadecane mildly increased with the increase of Pt/C loading (the yield of heptadecane increased from 7.4% to 20.0%, Figure 4c). Considering the different trend in the hydrothermal and dodecane reaction systems, there was likely another factor in hydrothermal reaction system, which inhabited the decarboxylation and aromatization reactions. The large amount of water released H+ and hydrogen bonding existing between oleic acid and water restricted the ionization of carboxyl groups, which may be another influencing factor. RCOOH↔RCOO-+H+ (small amount) H2O↔OH-+H+

(large amount)

In conclusion, the phenomenon that the decarboxylation and aromatization rates and yields of aromatics and heptadecane increased with an increase in Pt/C loading when

ACS Paragon Plus Environment

Page 14 of 45

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

solvents were added may be proof that competition for the active sites of the catalyst is an influencing factor. The restriction of carboxyl group ionization by the H+ and hydrogen bonding from water needs further research. 3.4 Effect of reaction time The conversion of oleic acid with respect to reaction time was examined. The same oleic acid loading (75 mg), Pt/C loading (15 mg), solvent loading (0.75 mL) and reaction temperature (350 °C) as before were used in the experiments in Section 3.4. The reaction time ranged from 1 to 6 h.

Figure 5 shows the effect of reaction time on the conversion of

oleic acid with different solvents (a: non-solvent

22

; b: dodecane; c: water). Figure 5a 22

shows oleic acid was quickly and completely converted within 15 min. On the contrary, the in-situ hydrogen transfer product of oleic acid (stearic acid) increased to 40% within 15 min and then decreased continuously with prolonged reaction time.

Finally, under

the condition of non-solvent, a heptadecane yield of 71% and aromatics yield of 19% were achieved within 80 min.

Figure 5b shows that in the dodecane reaction system,

stearic acid was transformed to heptadecane over time. When the reaction time increased to 3 h, stearic acid (intermediate) transformed completely, and the yield of heptadecane increased to the max value of 84%, while the yield of aromatics remained stable at 4%.

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

Figure 5c shows that in the hydrothermal reaction system oleic acid was hydrogenated to stearic acid very quickly.

When the reaction time increased to 4 h, stearic acid

transformed completely, and the yields of heptadecane and aromatics increased to 59% and 9%, respectively, and then the reaction remained stable.

In the dodecane and

hydrothermal reaction systems, the oleic acid and intermediate products could be completely consumed as the reaction time elapsed. Although the yield of aromatics was only 6% after 6 h in dodecane reaction system, the yield of heptadecane could rise to 83%, the mole balance of liquid sample was still as high as 89%. In the hydrothermal reaction system, the yields of heptadecane and aromatics were 61% and 7% after 6 h, the mole balance was about 75%, which might be caused by more coupling and polymerization reactions occurred in water.

Savage and co-workers

26

have reported that hydrogen

source added to the unsaturated fatty acids during the hydrothermal treatment is from hydrothermal gasification of both H2O and reactant with oleic acid as the reactant over Pt-based catalysts.

It might also be a reasonable explanation for why carbon balance is

low in the hydrothermal reaction system.

In the dodecane reaction system 23, hydrogen

transfer from dodecane can ensure the saturation of unsaturated fatty acid.

Reactant is

not used as a source of hydrogen, so carbon balance in the dodecane reaction system is

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

relative higher than that in the hydrothermal reaction system.

Unsaturated fatty acid is

easier to be hydrogenated in both hydrothermal and dodecane reaction system, as a result, aromatization reaction is suppressed compared with the non-solvent reaction system. Therefore, as the reaction time elapsed, fatty acids could still decarboxylate slowly in hydrothermal and dodecane reaction system, while the restriction for aromatization could not be eliminated by prolonging the reaction time. 3.5 Effect of solvent on the reaction of 1-heptadecene To prove that the restricted ionization of the carboxyl groups from the H+ of water and hydrogen bonding was an influencing factor on the reaction, experiments using 1-heptadecene as the reactant were carried out.

The same 1-heptadecene loading (75

mg), Pt/C loading (15 mg), solvent loading (0.75 mL), reaction temperature (350 °C) and reaction time (80 min) as before were used in the experiments of Section 3.5.

Figure 6

shows the solvents effect on the conversion of oleic acid and 1-heptadecene.

Using

oleic acid as a reactant, the yields of heptadecane and aromatics, were 71% and 19% respectively in the non-solvent system, 41% and 4% respectively in the dodecane system, 12% and 3% respectively in water system (Figure 6a).

It indicates that dodecane and

water inhibited the decarboxylation and aromatization of oleic acid compared to the data

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

under non-solvent condition.

However, the influence of water was much more obvious

than that of dodecane, suggesting that water possessed more influence factors than dodecane.

Using 1-heptadecene as a reactant, the yields of heptadecane and aromatics

were 65% and 13% respectively in the non-solvent system, 38% and 3% respectively in the dodecane system, 37% and 4% respectively in the water system (Figure 6b). Namely, the dodecane and water had the same influence on the conversion of 1-heptadecene compared with non-solvent reaction system, indicating that the influence way of water was the same as dodecane (competing with the catalyst) in the reaction system of 1-heptadecene.

The difference between oleic acid and 1-heptadecene is the

carboxyl group, whose ionization can be restricted by H+ released from water, and can also be used to form the hydrogen bonding.

The different experiment results in the

reaction system of oleic acid and 1-heptadecene indicate that in addition to the similar influencing factor to that in the dodecane system, H+ released from water and hydrogen bonding in the hydrothermal reaction system, inhibiting the ionization of carboxyl groups, was the key influence factor for the slower decarboxylation and aromatization rates under hydrothermal conditions.

ACS Paragon Plus Environment

Page 18 of 45

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

3.6 Reusability of Catalyst The reusability of catalyst were examined in the different solvent systems. The same oleic acid loading (75 mg), Pt/C loading (15 mg), solvent loading (0.75 mL), reaction temperature (350 °C) and reaction time (80 min) as before were used in the experiments in Section 3.6. Figure 7 shows the reaction results of oleic acid over spent Pt/C with different solvents (a: non-solvent; b: dodecane; c: water). In the non-solvent system, the yields of heptadecane decreased from 69% to 18% when Pt/C was used for the third time, while the intermediates 8-heptadecene and stearic acid increased from 2% to 12% and 0% to 43% respectively.

In the dodecane system, the yields of heptadecane decreased

from 41% to 15% when Pt/C was used for the third time, while the intermediates 8-heptadecene and stearic acid increased from 1% to 8% and 30% to 50% respectively. These results indicate that the activity of catalyst for decarboxylation reaction decreased after use, which might be caused by the depletion of Pt from the catalyst, carbon deposition or the adsorption of products on the catalyst.

While in the hydrothermal

system, the yields of stearic acid (59.7% to 68.1%) and heptadecane (12.5% to 10.0%) were relatively stable when Pt/C was used for the third time. The possible reason was suspected as follows.

a) In the hydrothermal system, decarboxylation rate was slow

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 20 of 45

owing to the H+ released from water and hydrogen bonding between the water and oleic acid.

The changes in active site of the catalyst after use did not significantly affect the

efficiency of decarboxylation, consistent with the results in Figures 4b and 4c.

As the

Pt/C loading increased from 5 mg to 10 mg, the yield of heptadecane increased remarkably from 10 % to 35% in dodecane system, but it only rose from 7% to 11% in hydrothermal system.

As a result, the decarboxylation rates in the water did not change

remarkably and yield of heptadecane varied from 12.5% to 10.0% when Pt/C was used for the third time. b) The deactivation of catalyst in the hydrothermal system was not as serious as those in the other two systems.

Characterizations (ICP-MS, BET and TGA)

results were used for explaining relative slighter deactivation of catalyst in the hydrothermal system. In order to prove the Pt leaching, the liquid sample after reaction was rinsed and diluted, and the metal in the solution was determined by the inductively coupled plasma mass spectrometry, X Series II (Thermo Fisher Scientific).

The loss of Pt during the

reactions under the conditions of non-solvent, dodecane and water was shown in Table 1. The Pt depletions in these solvent systems were all less than 0.0006%, suggesting the deactivation of Pt/C was not related to the depletion of Pt.

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 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

Table 2 shows the physical properties of Pt/C in reactions with fresh catalyst and catalysts recycled from (RF) non-solvent, dodecane and water reactions. Compared with the fresh Pt/C, Pt/C recycled from the non-solvent, dodecane and water reactions exhibited much smaller surface areas and volumes, and larger pore sizes. It’s noticeable that Pt/C recycled from the hydrothermal system had a slightly larger surface area compared to that recycled from the dodecane or non-solvent systems. Relative larger surface area may be another reason why the decarboxylation and aromatization rates decreased slighter after the third use when the solvent was water.

The N2 adsorption-desorption isotherms and

pore diameters of fresh catalyst and catalyst recycled from (RF) non-solvent, dodecane and water reactions are shown in Figures 8a and 8b, respectively. Adsorption-desorption isotherms show similar sharp hysteresis loops for all examined catalysts, and their pore diameter distributions were the same. The variation of the physical properties of Pt/C might be caused by the coking or the asorption of products.

Figure 9 shows the thermogravimetric analysis (TGA) results of

Pt/C recycled from the non-solvent, dodecane, and hydrothermal systems. The first peak at 165 °C represents the heat released from the combustion of coking or desorption of products, and the second peak at approximately 360 °C might represent the heat released

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

from the combustion of the carbon support of Pt/C.

Page 22 of 45

In order to figure out whether the

first peak at 165°C stands for the combustion of coking or the desorption of products, the TGA of fresh Pt/C (a) and fresh Pt/C wetted by products (b) were conducted, shown in Figure 10.

The sample of fresh Pt/C wetted by products was obtained as follows: fresh

Pt/C was wetted by the solution of products (heptadecane and undecylbenzene) and followed by rinsing 3 times (all the operation process was the same as sample treatment) with acetone.

Figure 10a (fresh Pt/C) shows that the peak at approximately 360 °C

represented the heat released from the combustion of the carbon support of Pt/C.

Figure

10b shows that there is no peak from 150 °C to 200 °C, meaning that recycle method can wash all the products off.

Therefore the first peak at 165 °C in Figure 9 should be

attributed to the combustion of coking rather than the desorption of products.

Figure 9

shows that the coking in the non-solvent, dodecane and hydrothermal system were 12%, 23%, 14% respectively.

The amount of coking in the dodecane system was the highest,

because the coking might be achieved from dodecane. The amount of coking in the non-solvent system was similar to that in the hydrothermal system, since the amount of organics were the same in these two systems, and water had no significant influence for the formation of coking.

The relative less coking was also a reason why the

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

decarboxylation rate decreased slightly in the hydrothermal system when Pt/C was used for the third time compared with that in the dodecane and non-solvent system. Overall, Pt leaching was not detected in these three solvent systems.

However,

coking and the decrease of surface area were observed in all solvent systems, especially in the dodecane system.

Therefore, decarboxylation rates in the non-solvent and

dodecane systems decreased when Pt/C was used for third time.

In the hydrothermal

system, relative slighter deactivation of catalyst and low decarboxylation rates resulted from the addition of water might be the reasons why decarboxylation results did not vary remarkably when Pt/C was used for third time.

4. .Conclusions Three solvent systems (solvent-free, water, and dodecane) were employed for the conversion of oleic acid over Pt/C at 350 °C to study the influence of solvent. Decarboxylation, hydrogen transfer, and aromatization were observed in these three reaction systems. Compared to the non-solvent reaction system, much slower decarboxylation and aromatization rates and fewer aromatic products were observed in the hydrothermal and dodecane reaction systems.

The decarboxylation and aromatization

rates and yield of aromatics decreased with increasing dodecane loading in the dodecane

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

reaction system, and the decarboxylation and aromatization rates and yield of aromatics significantly decreased in the hydrothermal reaction system.

The competition of

dodecane for the active sites of Pt/C is largely responsible for the slow decarboxylation and aromatization rates.

In addition to a similar influencing factor to that in the

dodecane system, in the hydrothermal reaction system, the H+ released from water along with hydrogen bonding inhibited the ionization of carboxyl groups and also played a large role in contributing to the slower decarboxylation and aromatization rates.

Acknowledgements This work was supported by the National Natural Science Foundation of China (No. 21436007, 21676243), Zhejiang Provincial Natural Science Foundation of China (No. LR17B060002, LZ14B060002), and the funding from the Boeing Company.

References 1.

Verma, D.; Kumar, R.; Rana, B. S.; Sinha, A. K., Aviation fuel production from lipids

by a single-step route using hierarchical mesoporous zeolites. Energy Environ. Sci. 2011, 4, (5), 1667-1671.

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

2.

Smith, B.; Greenwell, H. C.; Whiting, A., Catalytic upgrading of tri-glycerides and

fatty acids to transport biofuels. Energy Environ. Sci. 2009, 2, (3), 262-271. 3.

Gupta, K. K.; Rehman, A.; Sarviya, R. M., Bio-fuels for the gas turbine: A review.

Renewable Sustainable Energy Rev. 2010, 14, (9), 2946-2955. 4.

Biller, P.; Riley, R.; Ross, A. B., Catalytic hydrothermal processing of microalgae:

decomposition and upgrading of lipids. Bioresour. Technol. 2011, 102, (7), 4841-8. 5.

Davidson, C.; Newes, E.; Schwab, A.; Vimmerstedt, L., An overview of aviation fuel

markets for biofuels stakeholders. National Renewable Energy Laboratory, Retrieved September 2014, 5, 2014 6. Sacia, E. R.; Balakrishnan, M.; Deaner, M. H.; Goulas, K. A.; Toste, F. D.; Bell, A. T., Highly Selective Condensation of Biomass-Derived Methyl Ketones as a Source of Aviation Fuel. ChemSusChem 2015, 8, (10), 1726-1736. 7.

Šimáček, P.; Kubička, D.; Šebor, G.; Pospíšil, M., Hydroprocessed rapeseed oil as a

source of hydrocarbon-based biodiesel. Fuel 2009, 88, (3), 456-460. 8.

Meng, X.; Yang, J.; Xu, X.; Zhang, L.; Nie, Q.; Xian, M., Biodiesel production from

oleaginous microorganisms. Renewable Energy 2009, 34, (1), 1-5. 9.

Fu, J.; Shi, F.; Thompson, L. T.; Lu, X.; Savage, P. E., Activated Carbons for

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

Hydrothermal Decarboxylation of Fatty Acids. ACS Catal. 2011, 1, (3), 227-231. 10. Fu, J.; Lu, X.; Savage, P. E., Hydrothermal decarboxylation and hydrogenation of fatty acids over Pt/C. ChemSusChem 2011, 4, (4), 481-6. 11. Simakova, I.; Simakova, O.; Mäki-Arvela, P.; Murzin, D. Y., Decarboxylation of fatty acids over Pd supported on mesoporous carbon. Catal. Today 2010, 150, (1), 28-31. 12. Ping, E. W.; Pierson, J.; Wallace, R.; Miller, J. T.; Fuller, T. F.; Jones, C. W., On the nature of the deactivation of supported palladium nanoparticle catalysts in the decarboxylation of fatty acids. Appl. Catal., A 2011, 396, (1), 85-90. 13. Mäki-Arvela, P.; Kubickova, I.; Snåre, M.; Eränen, K.; Murzin, D. Y., Catalytic deoxygenation of fatty acids and their derivatives. Energy Fuels 2007, 21, (1), 30-41. 14. Peng, B.; Yuan, X.; Zhao, C.; Lercher, J. A., Stabilizing catalytic pathways via redundancy: selective reduction of microalgae oil to alkanes. J. Am. Chem. Soc. 2012, 134, (22), 9400-5. 15. Chen, H.; Wang, Q.; Zhang, X.; Wang, L., Quantitative conversion of triglycerides to hydrocarbons over hierarchical ZSM-5 catalyst. Appl. Catal., B 2015, 166-167, 327-334. 16. Wu, J.; Shi, J.; Fu, J.; Leidl, J. A.; Hou, Z.; Lu, X., Catalytic Decarboxylation of Fatty Acids to Aviation Fuels over Nickel Supported on Activated Carbon. Sci. Rep. 2016,

ACS Paragon Plus Environment

Page 26 of 45

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 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

6, 27820. 17. Yang, L.; Tate, K. L.; Jasinski, J. B.; Carreon, M. A., Decarboxylation of Oleic Acid to Heptadecane over Pt Supported on Zeolite 5A Beads. ACS Catal. 2015, 5, (11), 6497-6502. 18. Al Alwan, B.; Salley, S. O.; Ng, K. Y. S., Biofuels production from hydrothermal decarboxylation of oleic acid and soybean oil over Ni-based transition metal carbides supported on Al-SBA-15. Appl. Catal., A 2015, 498, 32-40. 19. Snåre, M.; Kubičková, I.; Mäki-Arvela, P.; Chichova, D.; Eränen, K.; Murzin, D. Y., Catalytic deoxygenation of unsaturated renewable feedstocks for production of diesel fuel hydrocarbons. Fuel 2008, 87, (6), 933-945. 20. Na, J.-G.; Yi, B. E.; Han, J. K.; Oh, Y.-K.; Park, J.-H.; Jung, T. S.; Han, S. S.; Yoon, H. C.; Kim, J.-N.; Lee, H.; Ko, C. H., Deoxygenation of microalgal oil into hydrocarbon with precious metal catalysts: Optimization of reaction conditions and supports. Energy 2012, 47, (1), 25-30. 21. 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. Catal. Sci. Technol. 2015, 5, (1), 380-388.

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

22. Tian, Q.; Qiao, K.; Zhou, F.; Chen, K.; Wang, T.; Fu, J.; Lu, X.; Ouyang, P., Direct Production of Aviation Fuel Range Hydrocarbons and Aromatics from Oleic Acid without an Added Hydrogen Donor. Energy Fuels 2016, 30, (9), 7291-7297. 23. Immer, J. G.; Kelly, M. J.; Lamb, H. H., Catalytic reaction pathways in liquid-phase deoxygenation of C18 free fatty acids. Appl. Catal., A 2010, 375, (1), 134-139. 24. Simakova, I.; Rozmysłowicz, B.; Simakova, O.; Mäki-Arvela, P.; Simakov, A.; Murzin, D. Y., Catalytic Deoxygenation of C18 Fatty Acids Over Mesoporous Pd/C Catalyst for Synthesis of Biofuels. Top. Catal. 2011, 54, (8-9), 460-466. 25. Fu, J.; Lu, X.; Savage, P. E., Catalytic hydrothermal deoxygenation of palmitic acid. Energy Environ. Sci. 2010, 3, (3), 311-317. 26. Yeh, T. M.; Hockstad, R. L.; Linic, S.; Savage, P. E., Hydrothermal decarboxylation of unsaturated fatty acids over PtSnx/C catalysts. Fuel 2015, 156, 219-224. 27. Mäki-Arvela, P.; Snåre, M.; Eränen, K.; Myllyoja, J.; Murzin, D. Y., Continuous decarboxylation of lauric acid over Pd/C catalyst. Fuel 2008, 87, (17-18), 3543-3549. 28. Vardon, D. R.; Sharma, B. K.; Jaramillo, H.; Kim, D.; Choe, J. K.; Ciesielski, P. N.; Strathmann, T. J., Hydrothermal catalytic processing of saturated and unsaturated fatty acids to hydrocarbons with glycerol for in situ hydrogen production. Green Chem. 2014,

ACS Paragon Plus Environment

Page 28 of 45

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

16, (3), 1507. 29. Shim, J. O.; Jeong, D. W.; Jang, W. J.; Jeon, K. W.; Kim, S. H.; Jeon, B. H.; Roh, H. S.; Na, J. G.; Oh, Y. K.; Han, S. S.; Ko, C. H., Optimization of unsupported CoMo catalysts for decarboxylation of oleic acid. Catal. Commun. 2015, 67, 16-20. 30. Kandel, K.; Anderegg, J. W.; Nelson, N. C.; Chaudhary, U.; Slowing, I. I., Supported iron nanoparticles for the hydrodeoxygenation of microalgal oil to green diesel. J. Catal. 2014, 314, 142-148. 31. Ng, J. B.; Shurvell, H., Application of factor analysis and band contour resolution techniques to the Raman spectra of acetic acid in aqueous solution. J. Phys. Chem. 1987, 91, (2), 496-500.

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

Table 1. Loss of Pt in the reactions with the condition of non-solvent, dodecane and water.

a

Solvent system

Pt depletiona

Non-solvent

0.00025%

Dodecane

0.00055%

Water

0.00030%

The mass ration of Pt lost in the liquid product and Pt added to the reactor

ACS Paragon Plus Environment

Page 30 of 45

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

Table 2. Physical properties of Pt/C in the reactions with fresh catalyst and catalysts recycled from (RF) non-solvent, dodecane, and water reactions.

a

b

Catalyst

SBET (m2/g)

Vtotal (cm3/g)a

Pore size (nm)b

Fresh

1,321.67

1.19

4.8

RF non-solvent

475.27

0.58

5.5

RF dodecane

492.85

0.65

5.5

RF water

546.03

0.61

5.7

Single point adsorption total pore volume of pores less than 193.4646 nm diameter at P/Po = 0.99

BJH Adsorption average pore diameter (4V/A)

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

Figure 1. A typical GC/FID chromatogram of the oleic acid in-situ hydrogen transfer, decarboxylation/decarbonylation and aromatization products under the condition of non-solvent.

ACS Paragon Plus Environment

Page 32 of 45

Page 33 of 45

80% Heptadecane Aromatics Stearic acid

60%

Yield

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

40%

20%

0% Non-solvent

Dodecane

Water

Figure 2. The yields of heptadecane, aromatics and stearic acid from oleic acid with different solvents over Pt/C. Reaction conditions: T = 350 °C, t = 80 min, oleic acid loading = 75 mg, Pt/C loading = 15 mg, and solvent loading = 0.75 mL.

ACS Paragon Plus Environment

Energy & Fuels

100% Heptadecane Aromatics Stearic acid

80%

Yield

60%

40%

20%

0% 0.00

0.25

0.50

0.75

1.00

Dodecane loading/mL

(a)

100% Heptadecane Aromatics Stearic acid

80%

Yield

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

60%

40%

20%

0% 0.00

0.25

0.50

0.75

1.00

Water loading/mL

(b) Figure 3. The effect of solvent (a: dodecane; b: water) loading on the conversion of oleic acid. Reaction conditions: T = 350 °C, t = 80 min, oleic acid loading = 75 mg, Pt/C loading = 15 mg, and solvent loading = 0-1 mL.

ACS Paragon Plus Environment

Page 35 of 45

75%

60%

Heptadecane Aromatics Stearic acid

Yield

45%

30%

15%

0% 5

10

15

20

Catalyst loading (mg)

(a)22

100%

80%

60% Yield

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

40%

Heptadecene Heptadecane Aromatics Stearic acid Balance

20%

0% 5

10

15

Catalyst loading (mg)

(b)

ACS Paragon Plus Environment

20

Energy & Fuels

100%

80%

60% Yield

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

Heptadecane Aromatics Oleic acid Stearic acid Balance

40%

20%

0% 5

10

15

20

Catalyst loading (mg)

(c) Figure 4. The effect of catalyst loading on the conversion of oleic acid with different solvents (a: non-solvent22; b: dodecane; c: water). Reaction conditions: T = 350 °C, t = 80 min, oleic acid loading = 75 mg, solvent loading = 0.75 mL, and Pt/C loading = 5-20 mg.

ACS Paragon Plus Environment

Page 37 of 45

90% 75% Heptedecene Heptadecane Aromatics Stearic acid Oleic acid

Yield

60% 45% 30% 15% 0% 0

10

20

30

40

50

60

70

80

Reaction time (min)

(a)22

90% 75% 60%

Yield

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

Heptadecane Aromatics Stearic acid

45% 30% 15% 0% 1

2

3

4

5

Reaction Time (h)

(b)

ACS Paragon Plus Environment

6

Energy & Fuels

100%

Other alkanes Heptadecane Aromatics Stearic acid

80%

60%

Yield

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

40%

20%

0% 1

2

3

4

5

6

Reaction time (h)

(c) Figure 5. The effect of reaction time on the conversion of oleic acid with different solvents (a: non-solvent22; b: dodecane; c: water). Reaction conditions: T = 350 °C, oleic acid loading = 75 mg, Pt/C loading = 15 mg, solvent loading = 0.75 mL, and t = 1-6 h.

ACS Paragon Plus Environment

Page 39 of 45

80%

80%

Heptadecane Aromatics Stearic acid

Heptadecane A romatics Heptadecene

60%

Yield

60%

Yield

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

40%

40%

20%

20%

0%

0% Non-solvent

Dodecane

Water

Non-solvent

Oleic acid as reactant

(a)

Dodecane

Water

1-heptadecene as reactant

(b)

Figure 6. The reaction results of oleic acid (a) and 1-heptadecene (b) with different solvents. Reaction conditions: T = 350 °C, t = 80 min, solvent loading = 0.75 mL, reactant loading = 75 mg, and Pt/C loading = 15 mg.

ACS Paragon Plus Environment

Energy & Fuels

80% 8-Heptadecene Heptadecane Aromatics Stearic acid

Yield

60%

40%

20%

0% 1st

2nd

3rd

(a)

80% 8-Heptadecene Heptadecane Aromatics Stearic acid

60%

Yield

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

40%

20%

0% 1st

2nd

3rd

(b)

ACS Paragon Plus Environment

Page 41 of 45

80%

8-Heptadecene Heptadecane Aromatics Stearic acid Oleic acid

60%

Yield

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

40%

20%

0% 1st

2nd

3rd

(c) Figure 7. The reaction results of oleic acid over recycled Pt/C with different solvents (a: non-solvent; b: dodecane; c: water). Reaction conditions: T = 350 °C, t = 80 min, solvent loading = 0.75 mL, oleic acid loading = 75 mg, and Pt/C loading = 15 mg.

ACS Paragon Plus Environment

Energy & Fuels

0.0 400

0.5

10

1.0

RF water

20

30

40

50

RF water

0.6 0.4

200

0.2

0 400

0.0 0.6

RF dodecane

RF dodecane

0.4

200

dV/dlog(D)

Vads (cm3/g STP)

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 42 of 45

0 400

RF non-solvent

0.2 0.0

RF non-solvent

0.6 0.4

200

0.2

0 800

0.0 1.5

Fresh

600

1.0

400

0.5

Fresh

200 0.0

0 0.0

0.5

P/P0

1.0

10

20

30

40

50

Pore diameter(nm)

(a)

(b)

Figure 8. N2 adsorption-desorption isotherms (a) and BJH pore diameter distributions (b) of Pt/C in the reactions with fresh catalyst and catalyst recycled from (RF) non-solvent, dodecane, and water reactions.

ACS Paragon Plus Environment

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 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

(a)

(b)

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

(c) Figure 9. Thermogravimetric analysis results of Pt/C in the reaction using catalyst recycled from non-solvent (a), dodecane (b), and water (c) reactions.

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 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

(a)

(b) Figure 10. Thermogravimetric analysis result of (a) fresh Pt/C and (b) fresh Pt/C wetted by the products (heptadecane and undecylbenzene) and followed by rinsing 3 times (all the processing steps were the same as samples)

ACS Paragon Plus Environment