Next Article in Journal
Passive Solar Photocatalytic Treatment of Emerging Contaminants in Water: A Field Study
Previous Article in Journal
Preparation and Analysis of Ni–Co Catalyst Use for Electricity Production and COD Reduction in Microbial Fuel Cells
Previous Article in Special Issue
In-Situ Catalytic Fast Pyrolysis of Pinecone over HY Catalysts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Catalytic Hydrodeoxygenation of Fast Pyrolysis Bio-Oil from Saccharina japonica Alga for Bio-Oil Upgrading

1
Department of Chemical Engineering, Kangwon National University, Samcheok, Gangwon-do 25913, Korea
2
Department of Chemical Engineering, Kyung Hee University, Giheung-gu, Yongin, Gyeonggi-do 17104, Korea
3
Department of Chemical and Materials Engineering, University of Kentucky, 4810 Alben Barkley Drive, Paducah, KY 42002, USA
4
Department of Chemical Engineering, Pukyong National University, 365 Sinseon-ro, Nam-gu, Busan 48513, Korea
*
Authors to whom correspondence should be addressed.
Catalysts 2019, 9(12), 1043; https://doi.org/10.3390/catal9121043
Submission received: 17 October 2019 / Revised: 27 November 2019 / Accepted: 3 December 2019 / Published: 8 December 2019
(This article belongs to the Special Issue Catalytic Fast Pyrolysis)

Abstract

:
Biomass conversion via pyrolysis has been regarded as a promising solution for bio-oil production. Compared to fossil fuels, however, the pyrolysis bio-oils from biomass are corrosive and unstable due to relatively high oxygen content. Thus, an upgrading of bio-oil is required to reduce O component while improving stability in order to use it directly as fuel sources or in industrial processes for synthesizing chemicals. The catalytic hydrodeoxygenation (HDO) is considered as one of the promising methods for upgrading pyrolysis bio-oil. In this research, the HDO was studied for various catalysts (HZSM-5, metal, and metal-phosphide catalysts) to improve the quality of bio-oil produced by fast pyrolysis of Saccharina japonica (SJ) in a fluidized-bed reactor. The HDO processing was carried out in an autoclave at 350 °C and different initial pressures (3, 6, and 15 bar). During HDO, the oxygen species in the bio-oil was removed primarily via formation of CO2 and H2O. Among the gases produced through HDO, CO2 was observed to be most abundant. The C/O ratio of produced bio-oil increased when CoMoP/γ-Al2O3, Co/γ-Al2O3, Fe/γ-Al2O3, or HZSM-5 was used. The Co/γ-Al2O3 resulted in higher HDO performance than other catalysts. The bio-oil upgraded with Co/γ-Al2O3 showed high HHV (34.41 MJ/kg). With the use of catalysts, the kerosene-diesel fraction (carbon number C12–C14) was increased from 36.17 to 38.62–48.92 wt.%.

1. Introduction

Biomass has attracted considerable attention as an alternative energy resource to replace nonrenewable fossil fuels. Biomass feedstocks including agricultural residues [1], forestry waste [2,3], plastic [4], aquatic plant [5,6], and sewage sludge [7], are inexpensive and can be further utilized to produce a variety of valuable chemicals and biomaterials as well as energy [5,8]. Among the various resources for renewable energy, macroalgae offers great promise as feedstock for biofuel production due to its higher growth rate even in wastewater and seawater, higher yield, and shorter harvesting cycle, compared to terrestrial biomass. Thermal conversion via pyrolysis has received a lot of attention as a promising technique to produce biofuel from macroalgae. Kim et al. [9,10] investigated the pyrolysis characteristics and kinetics of Saccharina japonica (SJ) alga and Sargassum sp. and found that the bio-oil produced at optimum conditions showed a higher heating value (HHV) of 31.9 MJ/kg and an O/C molar ratio of 0.16. Ly et al. [5] produced bio-oils with HHVs in the range of 24.75–28.72 MJ/kg by fast pyrolysis of SJ. By pretreating Enteromorpha clathrata alga with diluted HCl, Cao et al. [11] enhanced the yield of the bio-oil by 9.6%.
However, pyrolytic bio-oils contain a lot of oxygen-containing compounds, leading to a decrease in the bio-oil quality [3,12]. For this reason, in order for bio-oil to be directly used as transportation or boiler fuel, an upgrading process to lower the oxygen content in bio-oil is required. Catalytic pyrolysis is known as one of the most promising methods for upgrading bio-oil. In the presence of catalyst, the O-species in the bio-oils are released as forms of CO2, CO, and/or H2O, or converted directly to hydrocarbons at atmospheric pressure, resulting in improvement of bio-oil quality [12]. Ly et al. [3] also studied the catalytic pyrolysis of tulip tree (Liriodendron) in a fluidized-bed reactor with dolomite catalyst. They reported that oxygen content in the bio-oil was mostly removed as water via dehydration. In addition, gaseous products released together were found to have a high H2/CO ratio. The effect of HZSM-5 as a catalyst on the composition of pyrolyzed bio-oil from microalgae with high contents of nitrogen and oxygen was investigated by Lorenzetti et al. [13]. They found that the upgraded bio-oil contained high amounts of aromatic hydrocarbon, while contents of oxygen and nitrogen were relatively small. The catalytic deoxygenation over hydrotreating, known as catalytic hydrodeoxygenation (HDO), has been accepted as a potential method for efficient removal of oxygen in bio-oil. Due to high catalytic activity along with inexpensive material cost, alumina, zeolites, silica, and activated carbon have been commonly employed as supports for various HDO catalysts. Selecting 2-furyl methyl ketone (FMK) as a model compound, Ly et al. [14] obtained relatively high conversion (>83%) of FMK with Ni-based catalysts (Ni/γ-Al2O3). They also found that HDO was enhanced by addition of phosphorous, which influences the structural properties of active phase [14]. In particular, transition metal phosphide catalysts have been reported to show excellent activities on HDO [14,15,16]. The HDO processes with supported-metallic catalysts were investigated to upgrade pyrolytic bio-oils [17,18]. Cheng et al. [18] studied the HDO of bio-oil from pine sawdust with Fe-Co/SiO2 catalysts and found that the bimetallic Fe-Co/SiO2 catalysts (Fe/Co ratio of 1:1) showed better performance, as compared to monometallic catalysts (e.g., Fe/SiO2 and Co/SiO2).
In the present study, different catalysts including HZSM-5, metal (Fe and Co), and metal phosphide (Fe2P, CoP, and CoMoP) supported on γ-Al2O3 were examined for upgrading bio-oil obtained by pyrolysis of SJ. In addition, the feasibility of applying inexpensive catalysts to HDO processes was investigated, by testing in an autoclave at 350 °C and pressure varying from 3 to 15 bar. Finally, the compositions of the products and the catalysts were systematically analyzed using various techniques.

2. Results and Discussion

2.1. Catalysts and Sample Characterization

The X-ray diffractogram of catalysts was reported in Figure 1. The peaks at 2θ = 31.93°, 37.67°, 45.78°, and 66.35° were associated with γ-Al2O3 support. Two main peaks for HZSM-5 were displayed at 22.86° and 23.89° [19]. The peaks related to Fe phase in 10 wt.% Fe/γ-Al2O3 catalyst were observed at 2θ = 44.64° and 65.05°. The peaks of Fe3O4 appeared at 2θ values of 18.77°, 35.87°, 57.64°, and 62.91°. This result indicates that the Fe2O3 phase was transformed into Fe3O4 phase and further to Fe metallic phase. The peak corresponding to Fe2P could be observed at 38.4°. In the XRD patterns for Co-containing catalysts, the peaks of CoP phase appeared at 2θ = 31,62°, 35.23°, 40.12°, 48.19°, and 56.69°, Co at 42.39°, 51.24°, 54.01°, and 75.45°, and the peaks of spinel Co3O4 at 31.89°, 37.02°, 59.64°, and 65.86°. The peaks 25.94°, 36.98°, and 69.88° were attributed to MoO2, while the signals at 28.03°, 32.28°, and 43.20° for the spinel MoP. These XRD results were consistent with those reported in previous studies [14,15,16].
The BET surface area (SBET), the pore volume, and the average pore size of HZSM-5 were determined to be 132.49 m2/g, 0.14 cm3/g, and 6.70 nm, respectively. As presented in Table 1, the prepared catalysts (Co, CoP, Fe, Fe2P, and CoMoP on γ-Al2O3) showed the textual properties with specific surface area ranging from 184.82 to 211.08 m2/g, the pore volume 0.40–0.43 cm3/g, and the average pore size 7.05–8.09 nm. As reported in our prior study [9], the moisture and ash contents of SJ sample were 6.90 wt.% and 20.21 wt.%, respectively. Table 2 shows the characteristics of SJ and SJO. As shown in Table 2, the HHV of SJO was determined to be 26.10 MJ/kg, which was higher than that of woody biomass such as tulip tree (18.87 MJ/kg) [3] and Pinyon pine (18.94 MJ/kg) [2].

2.2. Thermogravimetric Analysis of Biomass Samples

The thermogravimetric analysis (TGA) combined with differential thermogravimetric analysis (DTG) is a high-precision technique to understand the characteristic of thermal decomposition of solid materials. The TGA enables to monitor the degree of conversion with temperature, while the differential rate of conversion (dX/dt) vs. temperature is expressed by DTG. Both TGA and DTG curves for original SJ and SJO at a heating rate of 20 °C/min are illustrated in Figure 2. Starting at −40 °C, most of SJO conversion was obtained below 500 °C, which is ascribed to the decomposition of organic components in the SJO. According to the simulated distillation by TGA [3,12,20], this decomposition temperature range (40–500 °C) corresponded to the boiling point of the organic products with carbon numbers from C5 to C38. It is also noteworthy that 80% of SJO was recovered within 350 °C, and the maximum differential rate of conversion was observed at 200.43 °C in the DTG curve (b).

2.3. Effect of Reaction Conditions on Product Distribution

The effect of initial pressure on the HDO of SJO was investigated. With increasing reactor pressure, the gas yield decreased, while the liquid yield increased. It is likely because increased pressure results in the suppression of volatiles. The moisture content in the oil decreased with increasing reaction pressure, whereas the content of organic matter increased. The HHV of the bio-oil was in the range of 33.09–33.99 MJ/kg, which was little influenced by the pressure. As compared to raw bio-oil, however, the HHV of the bio-oil upgraded by HDO increased because oxygen species are reduced via dehydration, decarboxylation, and decarbonylation. The HHV of the upgraded bio-oils was higher than that of pyrolysis bio-oils from biomass such as tulip tree (24.37 MJ/kg) [3] and S. japonica (26.1–28.27 MJ/kg) [9].
The effect of catalysts on the HDO of SJO was also investigated for HZSM-5 and various catalysts with 10 wt.% metal loading at 350 °C and the initial pressure of 15 bar. As shown in Table 3, the product distributions were different for the catalysts tested, which were likely due to different reaction pathways. When HDO of SJO was conducted without catalyst, the liquid yield decreased, whereas the gas yield increased, compared to catalytic HDO. The maximum liquid yield was 78.60 wt.%, observed for HZSM-5. The highest value of gas yield was 15.21 wt.%, achieved by HDO using Co/Al2O3. As a result, the addition of metal on γ-Al2O3 support (Co/γ-Al2O3 and Fe/γ-Al2O3) decreased the liquid and char yields, while the gas yield was increased. This might be due to the catalytic cracking reaction, which produced lower molecular weight compounds and non-condensable gas, followed by a series of deoxygenation reactions [12]. When HZSM-5 was applied, the liquid yield was decreased, but the char yield was increased, compared to HDO in the absence of the catalyst. An increase in the char yield might be due to the formation of coke on the surface of HZSM-5 by repolymerization and aromatization of aromatic compounds in liquid products [12].
When phosphorus was added into the metal catalysts, however, significant changes in the solid yield were observed. The solid yield of P-containing catalysts was higher than those of the P-free counterpart. The solid yields for Co/γ-Al2O3 and Fe/γ-Al2O3 were determined to be 15.21% and 14.39%, respectively. On the other hand, they were decreased to 8.8% and 7.92% with an addition of phosphorus (CoP/γ-Al2O3 and Fe2P/γ-Al2O3). This might be because the addition of phosphorus enhanced the acidity of catalyst, leading to the decrease of solid yield on active sites of catalysts. The HHVs of bio-oils through HDO were determined to be 34.41 MJ/kg (Co/γ-Al2O3) and 34.28 MJ/kg (Fe/γ-Al2O3), while for phosphide catalysts in the range of 28.87–33.94 MJ/kg.

2.4. Compositions of Gas Product

Table 3 also shows gaseous product distributions by non-catalytic and catalytic HDO of SJO. The gas compositions mostly consisted of CO, CO2, CH4, and other hydrocarbon gases and varied depending on the reaction conditions. The CH4 was formed by cracking and hydrocracking of alkyl groups during the upgrading process. Without a catalyst, CH4 content in gas products increased from 7.61 to 10.32 mol% with increasing pressure from 3 to 15 bar. This was probably due to the increase in the demethylation reaction rate. The demethylation was favored in HDO with Co/γ-Al2O3 (10.24 mol%) and Fe/γ-Al2O3 (12.05 mol%) catalysts. In this study, CO2 was identified as a major component among produced gases (82.19–90.64 mol%), indicating that the organic compounds in raw SJO underwent decarboxylation [12]. For the HDO with catalysts, on the other hand, the decarboxylation reaction was more predominant than decarbonylation. Particularly for Fe2P/γ-Al2O3 and CoMoP/γ-Al2O3 catalysts, high selectivity to carbon dioxide was observed, which is likely to be attributed to promoted dehydration and decarboxylation by addition of phosphorus [21,22].

2.5. Bio-Oil Analysis

The atomic C/O ratio of bio-oils obtained by HDO was calculated based on the result of the elemental analysis of the bio-oil. By performing catalytic HDO, the C/O ratio of bio-oil increased from 4.12 (without catalyst) to 4.75 (HZSM-5), 6.34 (Co/γ-Al2O3), 5.67 (Fe/γ-Al2O3), and 6.75 (CoMoP/γ-Al2O3), while decreasing to 3.77 and 3.71 for CoP/Al2O3 and Fe2P/γ-Al2O3, respectively. The decrease in O content after HDO reaction was due to the removal of oxygenates from the gas (CO, CO2) or H2O via deoxygenation and dehydration reactions. The carbon and oxygen balances of the reactant and products were shown in Table 4.
The pyrolysis bio-oil is known as a complex mixture consisting of hundreds of components with a wide range of molecular weight. Table 5 shows the GC–MS analysis data of the bio-oils produced by HDO reaction at 15 bar and 350 °C with/without catalysts (based on peak area %). The bio-oil mainly contained components such as fatty acids, dianhydromannitol, isosorbide, 2-furyl methyl ketone (2-FMK), and derivatives of ketones. It can be observed that HZSM-5 was effective in the conversion of dianhydromannitol and isosorbide and the production of aromatic compounds such as derivatives of pyrazine, pyridinamine, and indole. This result was in good agreement with other literature [12]. An increase in the alkane selectivity, resulting in a decrease in the selectivity to fatty acids, might be explained by decarboxylation of fatty acids by catalytic activity of metal catalysts (Fe/Al2O3 and Fe2P/Al2O3) [23]. As shown in Table 4, the selectivity of ketone derivatives in the produced bio-oils after HDO, especially 2-FMK, was found to be lower than that of raw SJO. This result was also in good agreement with our previous study on HDO of bio-oil model compounds [14,15,16].
The simulated distillation using TGA, which is based on the boiling point of a carbonaceous liquid, has been conducted for the SJ bio-oil [3,12]. As shown in Figure 3, the carbon number distribution of bio-oil could be classified into three groups such as C5–C11, C12–C18, and C20–C38, corresponding to gasoline, kerosene-diesel, and heavy oil fractions, respectively. The distribution of these fractions in the SJO (without catalyst) was 36.24 (C5–C11), 36.17 (C12–C18), and 27.38 wt.% (C20–C38), respectively. However, there were significant changes in the carbon number distribution for the bio-oils by catalytic HDO. The fractions of C5–C11, C12–C18, and C20–C38 for bio-oil (organic phase) upgraded by HDO were in the range of 30.7–44.81, 38.62–48.92, and 14.16–28.64 wt.%, respectively.

3. Material and Methods

3.1. Sample and Catalysts Preparation

The SJ biomass used in this study was provided by the Cleaner Production Institute of Pukyong National University. The samples were subjected to drying at 105 °C overnight to achieve equilibrium moisture before the experiments. The bio-oils (organic phase) were obtained by fast pyrolysis at 450 °C and the HDO was proceeded at the fluidization velocity of 4.0 × Umf.
Following the ASTM standard method (ASTM E 1756-1 and ASTM E 1755-01), the ultimate and proximate analyses of biomass samples and bio-oil were conducted. The thermal decomposition of the bio-oil was analyzed using thermogravimetric analyzer (TGA N-1000, SINCO) under a nitrogen flow rate of 20 mL/min, from room temperature up to 700 °C at a heating rate 10 °C/min.
All catalysts tested in this study were ground and sieved to 80–100 mesh (150–180 μm). The commercial HZSM-5 catalyst was provided by Hyundai Petroleum Chem. Co. (South Korea) and was calcined at 550 °C for 5 h before use [24,25,26,27,28]. Alumina-supported monometallic (Co/γ-Al2O3 and Fe/γ-Al2O3) and metal phosphorus (Co/γ-Al2O3 and Fe/γ-Al2O3) catalysts were synthesized by impregnation method with 10 wt.% metal loading and a molar ratio of phosphorus (P) to metal (M) 1:1. Before the catalysts were used in the HDO process, they were calcined at 600 °C for 3 h and were further treated in the H2 environment at atmospheric pressure to reduce metal oxides or metal/metal-oxide phosphate to metal or metal phosphide, respectively. More details on catalysts preparation procedure and characterization methods are described in our prior work [15,16]. The specific surface area of the catalyst was determined using the multipoint Brunauer–Emmett–Teller (BET). Powder X-ray diffraction (XRD, MAC-18XHF, Rigaku, Japan) was also employed to understand the crystallographic structure of the catalysts.

3.2. Experimental Setup and Analytical Method

Upgrading of Saccharina japonica bio-oil (SJO) was carried out in an autoclave reactor. As shown in Figure 4, the system consists of a salt bath, a temperature controller, a mechanical stirrer, and a reactor with an inner volume of 100 mL. A molten salt mixture as heat transfer fluid was prepared from a eutectic salt of KNO3 (59 wt.%) and Ca(NO3)2 (41 wt.%) [6,10]. The experiments were conducted at a fixed temperature of 350 °C under different initial pressures from 3 to 15 bar using hydrogen. With the autoclave submerged in the molten salt bath, a ratio of catalyst to bio-oil of 1:10 (i.e., 1 g of catalyst with 10 g bio-oil sample) was used in the catalytic experiments. The residence time of the reactant in the salt bath was 60 min for each condition. After each run, the reactor was removed from the bath and cooled to room temperature.
The samples after HDO were collected to calculate product yields by determining the ratio of mass of the product to that of the biomass fed to the system. The gas yield was determined by measuring the difference in the weight of the reactor before and after the reaction. To calculate the solid and liquid yields, liquid and solid products were first separated by solvent extraction with acetone using a micro filter paper (pore size: 0.45 μm). Then, the solid yield was calculated by weighing the solid and filter after drying, while the liquid yield was given by difference. For all the calculations presented, each data point was an average of more than two experiments.
The elemental compositions of the upgraded bio-oils were characterized by Flash EA1112, CE Instrument [3,9,10]. The moisture content was measured by a Karl-Fischer (CA-200, Mitsubishi, Seoul, South Korea). The gas compositions were identified by the gas chromatography (YL 6500GC) equipped with dual detectors, a flame ionization detector (FID) using Porapak N column to identify hydrocarbon gases (C1–C4) and a thermal conductivity detector (TCD) using a Molecular sieve 13X column for H2, CO, CO2, and CH4. The FID was operated at 250 °C, using high-purity nitrogen (99.999%) as a carrier gas with flow rate 20 mL/min, while the TCD detector was held constant at 150 °C with a constant flow rate (20 mL/min) of argon (99.999% purity) as a carrier gas. Using helium carrier gas with a constant flow rate of 1.0 mL/min, a gas chromatograph/mass spectrometry (GC/MS, 7890A/5975C, Agilent, Seoul, South Korea) with a capillary column of HP-5MS (30 m × 0.25 mm × 0.25 µm) was applied to identify the compositions of bio-oils [3,5]. Starting at 40 °C, with a heating rate of 10 °C /min, the oven temperature increased to 280 °C and was maintained for 10 min. The temperature of injector was set constant at 280 °C and injection volume was 1 µL.

4. Conclusions

The HDO process of Saccharina japonica bio-oil was systematically investigated in an autoclave reactor. The HHVs of bio-oils upgraded by HDO were in the range of 33.74–33.99 MJ/kg in the absence of a catalyst. Although the liquid yield decreased, however, the quality of bio-oil increased by using metal catalysts in HDO of Saccharina japonica bio-oil. The HHVs of HDO bio-oils were increased to 34.41 MJ/kg by using Co/γ-Al2O3 catalyst but decreased with metal phosphide catalysts. The C/O ratio of HDO bio-oil with CoMoP/γ-Al2O3, Co/γ-Al2O3, Fe/γ-Al2O3, and HZSM-5 were higher than that of raw Saccharina japonica bio-oil and HDO bio-oil with CoP/γ-Al2O3 or Fe2P/γ-Al2O3 catalyst. Metal phosphide catalysts were likely to promote the decarboxylation, while metal catalyst elevating the demethylation reactions. The carbon number distribution of bio-oil was mainly distributed in the range C5–C11 and C12–C18 fractions. Our results revealed the feasibility of upgrading SJO to high-quality bio-oil using catalysts, and this upgraded bio-oil could be further used as a great source for manufacturing alternative bio-fuels and/or valuable chemicals.

Author Contributions

Conceptualization, H.V.L., S.-S.K. and J.K.; methodology, H.V.L. and S.-S.K.; formal analysis and investigation, H.V.L., J.H.C. and S.-S.K.; resources, J.H.C. and H.C.W.; writing—original draft preparation, H.V.L. and S.-S.K.; writing—review and editing, H.V.L., S.-S.K., H.T.H. and J.K.; supervision, S.-S.K. and J.K.; project administration, H.C.W. and S.-S.K.

Funding

This research received no external funding.

Acknowledgments

This research was supported by the Basic Science Research Program through the National Research Foundation of Korea (NRF), funded by the Ministry of Education, Science and Technology (NRF-2017R1E1A1A01074282).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zhang, S.; Dong, Q.; Zhang, L.; Xiong, Y. Effects of water washing and torrefaction on the pyrolysis behavior and kinetics of rice husk through TGA and Py-GC/MS. Bioresour. Technol. 2016, 199, 352–361. [Google Scholar] [CrossRef] [PubMed]
  2. Kim, S.-S.; Shenoy, A.; Agblevor, F. Themogravimetric and kinetic study of Pinyon pine in the various gases. Bioresour. Technol. 2014, 156, 297–302. [Google Scholar] [CrossRef] [PubMed]
  3. Ly, H.V.; Limm, D.-H.; Simm, J.W.; Kim, S.-S.; Kim, J. Catalytic pyrolysis of tulip tree (Liriodendron) in bubbling fluidized-bed reactor for upgrading bio-oil using dolomite catalyst. Energy 2018, 162, 564–575. [Google Scholar] [CrossRef]
  4. Honus, S.; Kumagai, S.; Molnar, V.; Fedorko, G.; Yoshioka, T. Pyrolysis gases produced from individual and mixed PE, PP, PS, PVC, and PET—Part II: Fuel characteristics. Fuel 2018, 221, 361–373. [Google Scholar] [CrossRef]
  5. Ly, H.V.; Kim, S.-S.; Woo, H.C.; Choi, J.H.; Suh, D.J.; Kim, J. Fast pyrolysis of macroalga Saccharina japonica in a bubbling fluidized-bed reactor for bio-oil production. Energy 2015, 93, 1436–1446. [Google Scholar] [CrossRef]
  6. Ly, H.V.; Kim, S.-S.; Kim, J.; Choi, J.H.; Woo, H.C. Effect of acid washing on pyrolysis of Cladophora socialis alga in microtubing reactor. Energy Convers. Manag. 2015, 106, 260–267. [Google Scholar] [CrossRef]
  7. Supaporn, P.; Ly, H.V.; Kim, S.-S.; Yeom, S.H. Bio-oil production using residual sewage sludge after lipid and carbohydrate extraction. Environ. Eng. Res. 2019, 24, 202–210. [Google Scholar] [CrossRef] [Green Version]
  8. Guedes, R.E.; Luna, A.S.; Torres, A.R. Operating parameters for bio-oil production in biomass pyrolysis: A review. J. Anal. Appl. Pyrol. 2018, 129, 134–149. [Google Scholar] [CrossRef]
  9. Kim, S.-S.; Ly, H.V.; Choi, J.H.; Kim, J.; Woo, H.C. Pyrolysis characteristics and kinetics of the alga Saccharina japonica. Bioresour. Technol. 2012, 123, 445–451. [Google Scholar] [CrossRef]
  10. Kim, S.-S.; Ly, H.V.; Kim, J.; Choi, J.H.; Woo, H.C. Thermogravimetric characteristics and pyrolysis kinetics of Alga Sagarssum sp. biomass. Bioresour. Technol. 2013, 139, 242–248. [Google Scholar] [CrossRef]
  11. Cao, B.; Wang, S.; Hu, Y.; Abomohra, A.E.-F.; Qian, L.; He, Z.; Wang, Q.; Uzoejinwa, B.B.; Esakkimuthu, S. Effect of washing with diluted acids on Enteromorpha clathrata pyrolysis products: Towards enhanced bio-oil from seaweeds. Renew. Energy 2019, 138, 29–38. [Google Scholar] [CrossRef]
  12. Ly, H.V.; Choi, J.H.; Woo, H.C.; Kim, S.-S.; Kim, J. Upgrading bio-oil by catalytic fast pyrolysis of acid-washed Saccharina japonica alga in a fluidized-bed reactor. Renew. Energy 2019, 133, 11–22. [Google Scholar] [CrossRef]
  13. Lorenzetti, C.; Conti, R.; Fabbri, D.; Yanik, J. A comparative study on the catalytic effect of H-ZSM5 on upgrading of pyrolysis vapors derived from lignocellulosic and proteinaceous biomass. Fuel 2016, 166, 446–452. [Google Scholar] [CrossRef]
  14. Ly, H.V.; Im, K.; Go, Y.; Galiwango, E.; Kim, S.-S.; Kim, J.; Choi, J.H.; Woo, H.C. Spray pyrolysis synthesis of γ-Al2O3 supported metal and metal phosphide catalysts and their activity in the hydrodeoxygenation of a bio-oil model compound. Energy Convers. Manag. 2016, 127, 545–553. [Google Scholar] [CrossRef]
  15. Le, T.A.; Ly, H.V.; Kim, J.; Kim, S.-S.; Choi, J.H.; Woo, H.C.; Othman, M.R. Hydrodeoxygenation of 2-furyl methyl ketone as a model conpound in bio-oil from pyrolysis of Saccharaina japonica alga in fixe-bed reactor. Chem. Eng. J. 2014, 105, 157–163. [Google Scholar] [CrossRef]
  16. Ly, H.V.; Galiwango, E.; Kim, S.-S.; Kim, J.; Choi, J.H.; Woo, H.C.; Othman, M.R. Hydrodeoxygenation of 2-furyl methyl ketone as a model compound of algal Saccharina Japonica bio-oil using iron phosphide catalyst. Chem. Eng. J. 2017, 317, 302–308. [Google Scholar] [CrossRef]
  17. Cheng, S.; Wei, L.; Zhao, X.; Julson, J. Application, deactivation, and regeneration of heterogeneous catalysts in bio-oil upgrading. Catalysts 2016, 6, 195. [Google Scholar] [CrossRef]
  18. Cheng, S.; Wei, L.; Julson, J.; Rabnawaz, M. Upgrading pyrolysis bio-oil through hydrodeoxygenation (HDO) using non-sulfided Fe-Co/SiO2 catalyst. Energy Convers. Manag. 2017, 150, 331–342. [Google Scholar] [CrossRef]
  19. Sun, J.; Liu, N.; Zhai, S.; Xiao, Z.; An, Q.; Huan, D. Gold-titania/protonated zeolite nanocomposite photocatalysts for methyl orange degradation under ultraviolet and visible irradiation. Mater. Sci. Semicond. Process. 2014, 25, 286–293. [Google Scholar] [CrossRef]
  20. Kim, S.-S.; Chun, B.H.; Kim, S.H. Non-isothermal pyrolysis of waste automobile lubricating oil in a stirred batch reactor. Chem. Eng. J. 2003, 93, 225–231. [Google Scholar] [CrossRef]
  21. Zepeda, T.A.; Infantes-Molina, A.; Diaz de Leon, J.N.; Fuentes, S.; Alonso-Nunez, G.; Torres-Otanez, G.; Pawelec, B. Hydrodesulfurization enhancement of heavy and light S-hydrocarbons on NiMo/HMS catalysts modified with Al and P. Appl. Catal. A Gen. 2014, 484, 108–121. [Google Scholar] [CrossRef]
  22. Prieur, B.; Meub, M.; Wittemann, M.; Klein, R.; Bellayer, S.; Fontaine, G. Phosphorylation of lignin: Characterization and investigation of the thermal decomposition. RSC Adv. 2017, 7, 16866–16877. [Google Scholar] [CrossRef] [Green Version]
  23. Ngo, T.-A.; Kim, J.; Kim, S.-S. Fast pyrolysis of spent coffee waste and oak wood chips in a micro-tubular reactor. Energy Source Part A 2015, 37, 1186–1194. [Google Scholar] [CrossRef]
  24. Yathavan, B.K.; Agblevor, F.A. Catalytic pyrolysis of Pinyon−Juniper using red mud and HZSM-5. Energy Fuels 2013, 27, 6858–6865. [Google Scholar] [CrossRef]
  25. Suchithra, T.-G.; Adhikari, S.; Chattanathan, S.A.; Gupta, R.B. Catalytic pyrolysis of green algae for hydrocarbon production using H+ZSM-5 catalyst. Bioresour. Technol. 2012, 118, 150–157. [Google Scholar] [CrossRef]
  26. Lee, H.; Kim, Y.M.; Lee, I.G.; Jeon, J.K.; Jung, S.C.; Chung, J.D.; Choi, W.G.; Park, Y.K. Recent advances in the catalytic hydrodeoxygenation of bio-oil. Korean J. Chem. Eng. 2016, 33, 3299–3315. [Google Scholar] [CrossRef]
  27. Lee, Y.J.; Shafaghat, H.; Kim, J.K.; Jeon, J.K.; Jung, S.C.; Lee, I.G.; Park, Y.K. Upgrading of pyrolysis bio-oil using WO3/ZrO2 and Amberlyst catalysts: Evaluation of acid number and viscosity. Korean J. Chem. Eng. 2017, 34, 2180–2187. [Google Scholar] [CrossRef]
  28. Kim, H.N.; Shafaghat, H.; Kim, J.K.; Kang, B.S.; Jeon, J.K.; Jung, S.C.; Lee, I.G.; Park, Y.K. Stabilization of bio-oil over a low cost dolomite catalyst. Korean J. Chem. Eng. 2018, 35, 922–925. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of (a) HZSM-5 and γ-Al2O3-supported catalyst with (b) 10 wt.% Co, (c) 10 wt.% CoP, (d) 10 wt.% CoMoP, (e) 10 wt.% Fe, and (f) 10 wt.% Fe2P.
Figure 1. XRD patterns of (a) HZSM-5 and γ-Al2O3-supported catalyst with (b) 10 wt.% Co, (c) 10 wt.% CoP, (d) 10 wt.% CoMoP, (e) 10 wt.% Fe, and (f) 10 wt.% Fe2P.
Catalysts 09 01043 g001
Figure 2. TGA and differential thermogravimetric analysis (DTG) of weight loss curves at a heating rate of 20 °C/min: (a) Saccharina japonica (SJ) macroalgae and (b) Saccharina japonica bio-oil (SJO).
Figure 2. TGA and differential thermogravimetric analysis (DTG) of weight loss curves at a heating rate of 20 °C/min: (a) Saccharina japonica (SJ) macroalgae and (b) Saccharina japonica bio-oil (SJO).
Catalysts 09 01043 g002
Figure 3. Carbon number distribution of bio-oil produced from HDO of Saccharina japonica bio-oil in an autoclave at 350 °C under pressure of 15 bar using different catalysts.
Figure 3. Carbon number distribution of bio-oil produced from HDO of Saccharina japonica bio-oil in an autoclave at 350 °C under pressure of 15 bar using different catalysts.
Catalysts 09 01043 g003
Figure 4. Schematic diagram of experimental apparatus for hydrodeoxygenation (HDO) of Saccharina japonica bio-oil.
Figure 4. Schematic diagram of experimental apparatus for hydrodeoxygenation (HDO) of Saccharina japonica bio-oil.
Catalysts 09 01043 g004
Table 1. Pore structure data of different catalysts.
Table 1. Pore structure data of different catalysts.
CatalystCalcination Temperature (°C)Surface Area (m2/g)Pore Volume (cm3/g)Average Pore Size (nm)
HZSM-5550132.490.046.7
10 wt.% Co/γ-Al2O3600184.820.407.79
10 wt.% Fe/γ-Al2O3600203.710.438.09
10 wt.% CoP/γ-Al2O3600209.870.437.05
10 wt.% Fe2P/γ-Al2O3600204.720.417.67
10 wt.% CoMoP/γ-Al2O3600211.080.427.56
Table 2. Characteristics of Saccharina japonica (SJ) and raw Saccharina japonica bio-oil (SJO).
Table 2. Characteristics of Saccharina japonica (SJ) and raw Saccharina japonica bio-oil (SJO).
Proximate Analysis (wt.%)Moisture aAsh bVolatile Matter cFixed Carbon cElemental Analysis d (wt.%)HHV (MJ/kg)
CHNO e
S.J [9]6.9020.2168.794.1032.896.170.9360.0112.11
Raw SJO2.14---61.968.022.2127.8127.45
a ASTM E1756, Standard test method for the determination of the total solids of biomass. b ASTM E1755, Standard test method for determination of ash content of biomass. c Calculating based on sample after drying. d On dry, ash free basis (for biomass material). e By difference.
Table 3. Product distribution of Saccharina japonica bio-oil for different conditions in an autoclave reactor at reaction time of 60 min.
Table 3. Product distribution of Saccharina japonica bio-oil for different conditions in an autoclave reactor at reaction time of 60 min.
Reaction Conditions350 °C,
3 Bar
350 °C,
6 Bar
350 °C, 15 Bar350 °C,
15 Bar, HZSM-5
350 °C,
15 Bar, Co/γ-Al2O3
350 °C,
15 Bar, Fe/γ-Al2O3
350 °C, 15 Bar, CoP/γ-Al2O3350 °C, 15 Bar, Fe2P/γ-Al2O3350 °C, 15 Bar,
CoMoP/γ-Al2O3
Product Yield (wt.%)Liquid77.24 ± 0.2178.09 ± 0.1880.17 ± 0.0878.60 ± 0.1174.28 ± 0.2073.63 ± 0.1675.14 ± 0.1172.76 ± 0.0768.58 ± 0.16
Moisture19.87 ± 0.0817.61 ± 0.1011.52 ± 0.0519.57 ± 0.1922.39 ± 0.2217.89 ± 0.1317.33 ± 0.0921.22 ± 0.1031.97 ± 0.25
Organic80.13 ± 0.0882.39 ± 0.1088.48 ± 0.0580.43 ± 0.1977.61 ± 0.2282.11 ± 0.1382.67 ± 0.0978.78 ± 0.1068.03 ± 0.25
Solid13.93 ± 0.2813.87 ± 0.1413.29 ± 0.1514.97 ± 0.2810.51 ± 0.3711.98 ± 0.2716.06 ± 0.1819.42 ± 0.2625.23 ± 0.40
Gas8.83 ± 0.078.04 ± 0.316.54 ± 0.226.43 ± 0.1715.21 ± 0.1814.39 ± 0.428.8 ± 0.307.82 ± 0.206.19 ± 0.23
Elemental Analysis (wt.%)C72.4471.4672.6073.1673.3973.4270.1663.8372.48
H8.468.398.598.388.548.428.438.228.58
N3.213.273.373.193.443.623.072.973.03
O15.8916.8815.4415.2614.6314.5418.3424.9815.91
HHV (MJ/kg) 33.7433.0933.9933.9934.4134.2832.5128.8733.94
Gas Selectivity (mol%)CH47.618.2510.32 9.0510.2412.059.27 5.22 5.17
C2H40.76 0.821.05 1.020.620.640.82 0.77 0.97
C2H61.69 1.752.10 2.081.791.771.43 1.35 1.10
C3H61.46 1.541.92 1.952.051.851.02 1.38 0.97
C3H80.54 0.560.62 0.830.870.600.37 0.470.29
CO0.18 0.491.67 0.830.720.901.36 0.17 1.83
CO287.76 86.5982.32 84.24 83.7182.1985.73 90.64 89.67
Table 4. Carbon and oxygen balance of the reactant and product during HDO reaction.
Table 4. Carbon and oxygen balance of the reactant and product during HDO reaction.
Reaction Conditions350 °C,
3 Bar
350 °C,
6 Bar
350 °C,
15 Bar
350 °C,
15 Bar,
HZSM-5
350 °C,
15 Bar,
Co/γ-Al2O3
350 °C,
15 Bar, Fe/γ-Al2O3
350 °C,
15 Bar, CoP/γ-Al2O3
350 °C,
15 Bar,
Fe2P/γ-Al2O3
350 °C,
15 Bar,
CoMoP/γ-Al2O3
Gas Product (g)C0.310.270.230.220.640.880.30.40.21
O0.540.510.450.410.630.910.510.430.33
Char Product (g)C1.3441.2611.1151.2611.0531.0111.4711.86 2.013
O0.060.090.490.160.110.090.10.080.06
Moisture Phase (g)C0.050.050.050.040.060.090.050.040.03
O1.361.220.821.371.481.171.161.371.95
Organic Phase (g)C4.43 4.55 4.74 4.60 4.39 4.19 4.31 3.82 3.86
O0.95 1.09 1.15 0.97 0.69 0.74 1.14 1.03 0.57
C/O Ratio 4.684.184.124.756.345.673.773.716.75
Table 5. Compounds identified by gas chromatograph/mass spectrometry (GC–MS) of bio-oil produced from HDO of Saccharina japonica bio-oil in an autoclave at 350 °C under pressure of 15 bar using various catalysts (results are based on peak area %).
Table 5. Compounds identified by gas chromatograph/mass spectrometry (GC–MS) of bio-oil produced from HDO of Saccharina japonica bio-oil in an autoclave at 350 °C under pressure of 15 bar using various catalysts (results are based on peak area %).
Composition of Bio-Oil Bio-Oil at 450 °C
4.0 × Umf
w/o CatalystCatalystsStructure
HZSM-5Co/γ-Al2O3CoP/γ-Al2O3CoMoP/γ-Al2O3Fe/γ-Al2O3Fe2P/γ-Al2O3
2-methyl-2-cyclopenten-1-one1.840.410.490.61 2.142.412.44 Catalysts 09 01043 i001
3-methyl-Butanal 3.53 3.66 Catalysts 09 01043 i002
2-Furyl methyl ketone12.76.535.622.455.227.74.366.11 Catalysts 09 01043 i003
3,4-dimethyl-2-cyclopentenone 1.64 1.32 Catalysts 09 01043 i004
2-hydroxy-3,4-dimethyl-2-Cyclopenten-1-one2.332.75 Catalysts 09 01043 i005
2-Hydroxy-3-ethyl-2-Cyclopenten-1-one1.251.82 Catalysts 09 01043 i006
2,3,6-Trimethylpyrazine 3.47 Catalysts 09 01043 i007
2,3-dimethyl-2-cyclopenten-1-one3.291.932.931.052.041.592.182.02 Catalysts 09 01043 i008
2,3,4-trimethyl-2-cyclopenten-1-one2.292.051.210.92.71 2.262.15 Catalysts 09 01043 i009
3-ethyl-2,5-dimethyl-Pyrazine4.231.760.67 Catalysts 09 01043 i010
2-methoxy-Phenol 0.26 3.140.77 Catalysts 09 01043 i011
2,4-Dimethyl phenol1.010.98 1.091.23 Catalysts 09 01043 i012
2,3-Dimethyl-5-ethyl-Pyrazine1.951.891.76 Catalysts 09 01043 i013
1-Acetyl-1-cyclohexene 2.61 Catalysts 09 01043 i014
4-Ethyl-2,5,6-Trimethylpyrimidine 2.42 Catalysts 09 01043 i015
Isosorbide3.482.66 4.0311.916.966.1218.52 Catalysts 09 01043 i016
Dianhydromannitol22.2620.733.3517.3319.6723.8220.6820.64 Catalysts 09 01043 i017
3-ethoxy-2-Pyridinamine 2.57 Catalysts 09 01043 i018
1,2-dimethoxy-Benzene 1.23 Catalysts 09 01043 i019
6,7-dihydro-2,5-dimethyl-5H-cyclopentapyrazine 2.63 Catalysts 09 01043 i020
Tetradecane 1.871.97 1.182.5 C14H30
n-Pentadecane 1.420.7 1.04 C15H32
n-Hexadecane 2.610.71 C16H34
2-Ethylhexyl 2-ethylhexanoate 2.83 Catalysts 09 01043 i021
Methyl n-tetradecanoate 1.091.782.491.781.971.02 Catalysts 09 01043 i022
Tetradecanoic acid 1.29 2.04 Catalysts 09 01043 i023
4-Benzylaniline 6.35 Catalysts 09 01043 i024
1,1-Diphenylhydrazine 2.65 Catalysts 09 01043 i025
3,7,11,15-tetramethyl-2-Hexadecene1.22.494.991.65 2.72 Catalysts 09 01043 i026
Methyl hexadecanoate0.460.871.281.443.422.32.641.21 Catalysts 09 01043 i027
Hexadecanoic acid5.163.441.471.912.231.09 Catalysts 09 01043 i028
9-Octadecenoic acid4.653.41 Catalysts 09 01043 i029
Methyl-9-octadecenoate 1.670.672.193.351.092.96 Catalysts 09 01043 i030
1-Methyl-2-phenyl-1H-indole 6.55 Catalysts 09 01043 i031
3-Nitrophthalic acid 37.4 Catalysts 09 01043 i032

Share and Cite

MDPI and ACS Style

Ly, H.V.; Kim, J.; Hwang, H.T.; Choi, J.H.; Woo, H.C.; Kim, S.-S. Catalytic Hydrodeoxygenation of Fast Pyrolysis Bio-Oil from Saccharina japonica Alga for Bio-Oil Upgrading. Catalysts 2019, 9, 1043. https://doi.org/10.3390/catal9121043

AMA Style

Ly HV, Kim J, Hwang HT, Choi JH, Woo HC, Kim S-S. Catalytic Hydrodeoxygenation of Fast Pyrolysis Bio-Oil from Saccharina japonica Alga for Bio-Oil Upgrading. Catalysts. 2019; 9(12):1043. https://doi.org/10.3390/catal9121043

Chicago/Turabian Style

Ly, Hoang Vu, Jinsoo Kim, Hyun Tae Hwang, Jae Hyung Choi, Hee Chul Woo, and Seung-Soo Kim. 2019. "Catalytic Hydrodeoxygenation of Fast Pyrolysis Bio-Oil from Saccharina japonica Alga for Bio-Oil Upgrading" Catalysts 9, no. 12: 1043. https://doi.org/10.3390/catal9121043

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop