Vous êtes sur la page 1sur 5

JOURNAL OF FUEL CHEMISTRY AND TECHNOLOGY Volume 38, Issue 5, October 2010 Online English edition of the Chinese

language journal Cite this article as: J Fuel Chem Technol, 2010, 38(5), 560 564 RESEARCH PAPER

Catalytic esterification of bio-oil by ion exchange resins


WANG Jin-jiang, CHANG Jie*, FAN Juan
Pulp & Paper Engineering State Key Laboratory, School of Chemistry and Chemical Engineering, South China University of Technology, Guangzhou 510641

Abstract: Upgradation of bio-oil before utilization is desirable to obtain high grade fuel because of its drawbacks like high viscosity, low heating value, poor stability and high corrosiveness. Organic acids in bio-oils can be converted to their corresponding esters by catalytic esterification and this greatly improved quality of bio-oils. We selected 732 and NKC-9 type ion exchanger resins as esterification catalysts for upgrading bio-oil. The catalytic activity was first investigated by model reaction. The esterification of bio-oil with methanol was conducted in a batch reactor. Acid numbers of upgraded bio-oil on 732 resin and NKC-9 resin were lowered by 88.54% and 85.95%, respectively, which represents the conversion of organic acids to neutral esters; the heating values increased by 32.26% and 31.64%, respectively; the H2O contents decreased by 27.74% and 30.87%, respectively; the densities were lowered by 21.77% for both and the viscosities fell by approximately 97%. A fixed bed reactor was used for continuous catalytic esterification of bio-oil on 732 resin, and the acid number remarkably decreased by 92.61%. The accelerated ageing test showed improvement of stability, and the aluminum strip corrosion test showed reduced corrosion rate of bio-oil after upgradation. Key words: bio-oil; ion exchange resin; catalytic esterification; stability; corrosiveness

Fast pyrolysis has been developed considerably as a front technology for efficient conversion of biomass into a liquid product known as bio-oil, which can be used as a potential substitute for fossil fuels. This technology is industrially feasible and has advantage in terms of reduced cost[13]. Great progress in fast pyrolysis of biomass has been achieved, and liquid product yield is more than 70% on dry feed[4,5]. However, bio-oil is not as good as fossil fuels and the drawbacks including high acidity, low heating value, high corrosiveness, high viscosity and poor stability limit its usage as high grade/transportation fuel. Consequently, upgrading of bio-oil before utilization is desirable to obtain a liquid product that can be used in a wider variety of applications. The current upgrading methods include catalytic hydrogenation, catalytic cracking and catalytic esterification[68]. The presence of a large amount of organic acids (formic acid, acetic acid and propionic acid) is the main reason for the strong acidity, high corrosiveness and poor stability of bio-oil[911]. In this work, we selected 732 and NKC-9 type ion exchange resins as catalysts[12,13] and tried to convert the organic acids in bio-oil into their corresponding neutral esters so as to improve the quality of bio-oil.

1.1

Materials and chemicals

The crude bio-oil produced by pyrolysis of wood chips in a circulating fluidized bed unit was provided by Devotion Group (Guangzhou, China). The capacity of bio-oil is 3000 ton/year, and the yield of bio-oil is close to 70%. 732 type cation exchange resin (reference standard: Amberlite IR-120) and NKC-9 type cation exchange resin (reference standard: Amberlyst 15) were commercially available and were pretreated before being used as esterification catalysts. The aluminum strips used in corrosion test and other reagents were of AR grade. 1.2 Model reaction

Experimental

The model reaction of esterification with methanol and acetic acid was used to evaluate the catalytic activities of 732 and NKC-9 resins. Acetic acid (57.5 mL) and methanol (81.5 mL) were added into a three-neck flask equipped with a thermometer and a reflux condenser. Five grams of resin catalyst was used. The reaction solution (0.5 mL) was sampled at a specific interval and measured quantitatively by NaOH standard solution titration. Using the titration method, the acetic acid conversion was calculated by the following equation: conversion = (1V/V0) 100%, in which V and V0

Received: 07-Jan-2010; Revised: 05-May-2010 * Corresponding author. Tel: +86 20 87112448, Fax: +86 20 87112448, E-mail: changjie@scut.edu.cn Foundation item: Supported by the Major State Basic Research Development Program of China (973 Program, 2010CB732205). Copyright 2010, Institute of Coal Chemistry, Chinese Academy of Sciences. Published by Elsevier Limited. All rights reserved.

WANG Jin-jiang et al. / Journal of Fuel Chemistry and Technology, 2010, 38(5): 560 564

are the volumes of the standard NaOH solution consumed in neutralizing the 0.5 mL solution sampled in the process and at the beginning of the reaction, respectively. Besides, model mixture, which comprised of several typical components as those in bio-oil, was prepared and used to explore the esterification activity of resin catalysts in a complex system similar to bio-oil. The ratio of components in model mixture was as follows[6,9]: m(phenol): m(dihydroxybenzene): m(acetic acid): m(furfural): m(H2O) = 19:11:15:30:25. Model mixture (30 mL) and methanol (60 mL) were added into the three-neck flask. The experiments were carried out at 50 C for 5 h, and 5 g catalyst was used. The pH value and acid number of model mixture were measured before and after the reaction. Acid number, which is expressed as milligrams of sodium hydroxide per gram of sample (mg NaOH/g) in this study, refers to the quantity of base required to titrate a sample to a specified end point. Bio-oil and its model mixture are usually a black or dark brown liquid, and therefore, it is difficult to choose a suitable visual indicator to signal the end point of the titration. Method of potentiometric titration identifies the end point by monitoring the greatest slope in the titration curve at the equivalence point, which is especially suitable for the titration of turbid or colored solutions. Metrohm 888 Titrando was used for the determination of acid numbers of bio-oil and its model mixture by potentiometric titration method. 1.3 Esterification of bio-oil

It was found that aluminum had poor resistance to corrosion of bio-oils[16,17], and the corrosion rate would be enhanced at higher temperature. The aluminum strips were machined into 3 cm 3.5 cm 0.1 mm and then immersed in bio-oils. At specific intervals, the strips were taken out and weighed and the weight loss was calculated. 1.4.3 Basic physical properties

The pH value was measured by a PHS-3C precision pH-meter from Shanghai Hongyi Instrumentation Co., Ltd. The dynamic viscosity was measured with SYD-265H Petroleum Products Kinematic Viscosity Tester (GB/T265-88) from Shanghai Changji Geological Instrument Co,.Ltd, and gross calorific value was measured using a WGR-1 calorimetric bomb (ASTM D4809). The water content was determined by the Karl Fischer titration (ASTM D1744, GB11146-89), which was performed by using Metrohm 788 Titrando. Density was measured by Capillary-stoppered pyknometer method.

Results and discussion

2.1 Catalytic activity of 732 resin and NKC-9 resin on model esterification The conversion of acetic acid over 732 and NKC-9 resins at 50 C or 70 C is shown in Fig. 2. At 70 C, conversion of acetic acid reached 84.72% within 120 min, with 732 resin as catalyst and no significant increase in the conversion after 60 min. The NKC-9 resin showed slightly lower activity than 732 resin, and acetic acid conversion rose with time remarkably and reached 79.74% in 120 min at 70 C. Conversion of acetic acid on 732 resin at 50 C could reach 74.12%. Both 732 and NKC-9 resins exhibited high activities for esterification of acetic acid.

The esterification of bio-oil was conducted in a batch reactor. Bio-oil and methanol were mixed in a volume ratio of 1:2 and were added into the three-neck flask equipped with a thermometer, a reflux condenser and a stirrer. The experiments were carried out at 50 C for 5 h, and the catalyst was used by 10 wt% of the bio-oil. A fixed bed reactor was also used for continuous esterification of bio-oil. Bio-oil and double volumes of methanol were mixed and pumped to the fixed bed ( 10 300 mm), which was kept at 50 C for 1 h. A schematic diagram of the fixed bed is shown in Fig. 1. 1.4 1.4.1 Bio-oil characterization Ageing test

The stability of bio-oils is determined by measuring the variations of viscosity under accelerated ageing conditions[14,15]. Thirty milliliter crude, diluted or upgraded bio-oil was placed in small sealed vials and heated at 80 C. Kinematic viscosity was measured at specific intervals. 1.4.2 Aluminum strip corrosion test

Fig. 1 Diagram of fixed bed reactor used in bio-oil catalytic esterification


1: Feedstock; 2: Pump; 3,5: Circulating water; 4: Fixed bed reactor; 6: Product

WANG Jin-jiang et al. / Journal of Fuel Chemistry and Technology, 2010, 38(5): 560 564

Fig. 2 Acetic acid conversion with time over different catalysts under different reaction temperatures
: 732, 70 C; : NKC-9, 70C; : 732, 50 C

Table 1 Characteristics of model mixture before and after catalytic esterification


Characteristic pH Acid number /mg NaOHg1 Model mixture 1.50 103.10 Before reaction 2.50 46.69 After reaction 732 2.40 4.39 NKC-9 2.68 4.46

Table 2 Characteristics of bio-oil before and after upgradation in batch reactor


Characteristics Acid number /mg NaOHg1 pH Density /kgm3
1

Crude bio-oil 53.86 2.51 1.24 16.62 15010.1 81.27

Before reaction 26.28 3.65 0.97 8.08 19321.0 2.48

Upgraded bio-oil 732 6.17 1.98 0.97 12.01 19852.8 2.46 NKC-9 7.57 2.70 0.97 11.49 19759.3 2.45

H2O content w /% Caloric value /kJkg Kinematic viscosity (at 40C)

/mm2s1

Table 1 shows that acid numbers of bio-oil model mixture changed significantly after esterification with methanol at 50 C for 5 h. Before reaction, acid number of model mixture decreased from 103.10 to 46.69 mg NaOH/g after dilution with double volumes of methanol. After esterificaiton on 732 and NKC-9 resins, the acid numbers of model mixture were lowered by 90.60% and 90.44%, respectively. This implies the high activity of resin catalyst for conversion of acetic acid even in the complex system similar to bio-oil. However, the change of pH values was irregular and became ambiguous. 2.2 Upgrading of bio-oil in batch reactor

was diluted with double volumes of methanol. After upgrading bio-oil on 732 resin and NKC-9 resin, acid numbers were greatly lowered by 88.54% and 85.95%, respectively. Organic acids in bio-oil were also proven to be esterified according to the results of Xiong[18] and Wang[19] in which also bio-oils were upgraded on the same resin catalysts. Besides, the H2O contents increased from 8.08% to 12.01%and 11.49%, respectively. This is logical as esterification would produce water. The densities of bio-oil were lowered to 0.97 kg/m3 for both for dilution with methanol, and the viscosities significantly reduced by approximately 97%. Compared to the caloric values of crude bio-oil, the caloric values of upgraded bio-oil increased by 32.26% and 31.64% after esterification on 732 and NKC-9 resins, respectively. However, the pH value of bio-oil was both lowered after esterification on 732 and NKC-9 resins, which showed inconformity with the change of acid numbers. According to Yao[20], the decrease in pH value might be due to the replacement of H+ ions of resins by the Ca2+, K+ and Na+ in bio-oils. If so, the resin catalyst would lose its catalytic activity because H+ ions are the main active site of the resin catalysts. Nevertheless, Fig. 3 indicates that the 732 resin recycled after bio-oil esterification reaction still remained highly active for esterification. The conversion of acetic acid reached 83.13% using recycled 732 resin within 2 h, which was close to that when using fresh 732 resin. From this point, the H+ ions would not be largely replaced. Several studies on upgrading bio-oil by catalytic esterification also found that pH value was lowered while the carboxylic acids in bio-oil were converted to their corresponding neutral esters[8,18,19]. Bio-oil is not a highly dispersed system, and according to Garcia-Perez [21], bio-oil is a mixture of multiphase structure. We consider that the phase structure of bio-oil was changed, and the undissolved strong acids in bio-oil were dissolved after upgrading, which would lead to the decrease of pH value. Furthermore, pH value corresponds to the concentration (activity, in fact) of H+ ions in a homogeneous solution. The complexity of bio-oil will lead to uncertainties of exact pH value during the measurements. The pH value may not reflect the true content of acids in bio-oil. Comparatively, acid number determination by acidbase titration can accurately quantify the total amount of acids in bio-oil and is more suitable for evaluating the esterification of bio-oil. 2.3 Upgrading of bio-oil in fixed bed reactor

The properties of bio-oil before and after upgradation are shown in Table 2. Before reaction, the acid number of bio-oil was lowered from 53.86 to 26.28 mg NaOH/g after the bio-oil

A fixed bed reactor filled with 732 resin was used for continuous esterification of bio-oil, and the properties of bio-oil are shown in Table 3. The acid number of bio-oil decreased by 84.86% after 1 h at 50 C; compared to the H2O content of crude bio-oil, the H2O content of upgraded bio-oil was reduced by 32.07%.

WANG Jin-jiang et al. / Journal of Fuel Chemistry and Technology, 2010, 38(5): 560 564

not show significant changes, and the viscosity varied between 1.91 and 1.98 mm2/s. The viscosity of bio-oil and methanol mixture before reaction was steady, and the viscosity increased from 1.97 to 2.04 mm2/s only after ageing for 110 h. Dilution with small amounts of alcohol is known to stabilize bio-oil, and because of the dilution effect by double volumes of methanol, the stability did not change significantly after esterification. 2.5 Corrosion property

Fig. 3 Conversion of acetic acid with time over fresh and recycled 732 resins
: Fresh 732 resin; : Recycled 732 resin

Table 3 Characteristics of bio-oil before and after upgraded in fixed-bed reactor


Characteristic Acid number /mg NaOHg1 pH value H2O content w/ % Crude bio-oil 53.86 2.51 16.62 Before reaction 26.28 3.65 8.09 Upgraded bio-oil 3.98 1.98 11.29

2.4

Stability

The kinematic viscosity of bio-oils ageing at 80 C is described in Fig. 4. Heating the crude bio-oil to 80 C totally altered its properties and the viscosity significantly increased with time. A serious phase separation was observed when bio-oil was heated, and the viscosity increased from 83.85 to 405.01 mm2/s after ageing for 96 h. The upgraded bio-oil did

Based on the weight loss, as shown in Fig. 5, upgraded bio-oil was less corrosive than the crude bio-oil. We confirmed that the property of uniform surface corrosion of bio-oil was improved after upgrading; however, several stains on the surface of aluminum strip were also observed after the corrosion of upgraded bio-oil. Whether pitting corrosion occurred on aluminum by upgraded bio-oil needs further study. We also found that the rate of weight loss of upgraded bio-oil is slightly faster than that of diluted bio-oil, which was not reacted. This may be due to an increase in moisture content after esterification. From the point of chemical equilibrium, esterification reaction will occur at room temperature with enough time even without catalysts. The catalytic process can accelerate the potential esterification and water production, and the esterified bio-oil can be treated further. After 35 days storage of aluminum strip in crude bio-oil, bio-oil diluted with methanol but not reacted, and upgraded bio-oil at room temperature, its weight loss was 1.023%, 1.175% and 0.819%, respectively. Upgraded bio-oil was less corrosive to aluminum than crude bio-oil and mixture of bio-oil and methanol before esterification.

Fig. 4 Variations of viscosity of bio-oil before and after reaction when aged at 80 C
: Crude bio-oil; : Bio-oil diluted with methanol before reaction; : Upgraded bio-oil

Fig. 5 Weight loss of aluminum strips corroded by crude bio-oil and upgraded bio-oil at 50 C
: Crude bio-oil; : Upgraded bio-oil; : Bio-oil diluted with methanol before reaction

WANG Jin-jiang et al. / Journal of Fuel Chemistry and Technology, 2010, 38(5): 560 564

Conclusions

upgrading of bio-oil catalyzed over solid acids. Journal of Fuel Chemistry and Technology, 2006, 34(6): 680683. [9] Branca C, Giudicianni P, Diblasi C. GC/MS Characterization of liquids generated from low-temperature pyrolysis of wood. Ind Eng Chem Res, 2003, 42(14): 31903202. [10] Mohan D, Pittman C U, Steele P H. Pyrolysis of wood/biomass for bio-oil: A critical review. Energy Fuels, 2006, 20(3): 848889. [11] Oasmaa A, Kuoppal A, Solantausta Y. Fast pyrolysis of forestry residue. 2. Physicochemical composition of product liquid. Energy Fuels, 2003, 17(2): 433443. [12] Cai H, Zhou B. Progress of ion exchange resins in organic catalysis. Chemical Industry and Engineering Progress, 2007, 26(3): 386391. [13] Li X, Wu D X, Xia C B, Li L F. Synthesis of ethyl acetate using cation exchange resin as catalyst. Fine Chemical Intermediates, 2006, 36(3): 5558. [14] Oasmaa A, Meier D. Norms and standards for fast pyrolysis liquids 1. Round robin test. J Anal Appl Pyrolysis, 2005, 73(2): 323334. [15] Lu Q, Yang X L, Zhu X F. Analysis on chemical and physical properties of bio-oil pyrolyzed from rice husk. J Anal Appl Pyrolysis, 2008, 82(2): 191198. [16] Lu Q, Zhang J, Zhu X F. Corrosion properties of bio-oil and its emulsions with diesel. Chin Sci Bull, 2008, 53(23): 37263734. [17] Darmstadt H, Gatcia-perez M, Adnot A, Chaala A, Kretschmer D, Roy C. Corrosion of metals by bio-oil obtained by vacuum pyrolysis of softwood bark residues. An X-ray photoelectron spectroscopy and auger electron spectroscopy study. Energy Fuels, 2004, 18(5): 12911301. [18] Xiong W M, Fu Y, Lai D M, Guo Q X. Upgrading of Bio-oil via Esterification Catalysted with Acidic Ion-exchange Resin. Chemical 17541758. [19] Wang Q, Yao Y, Wang S R, Luo Z Y, Cen K F. Experimental research on bio-oil catalytic esterification using cation exchange resins. Journal of Zhejiang University (Engineering Science), 2009, 43(5): 926930. [20] Yao Y. Separation and upgrading of bio-oil. Hangzhou: Zhengjiang University, 2008. [21] Garcia-perez M, Chaala A, Pakdel H, Kretschmer D, Rodrigue D, Roy C. Multiphase structure of bio-oils. Energy Fuels, 2006, 20(1): 364375. Journal of Chinese University, 2009, 30(9):

732 and NKC-9 type ion exchange resins were selected as esterification catalysts in model reaction of acetic acid and methanol. Both 732 and NKC-9 resins exhibited high activities for esterification of acetic acid. After bio-oil upgradation on 732 and NKC-9 in a batch reactor, the acid number of bio-oil decreased by 88.54% and 85.95%, respectively, which represents the large conversion of organic acids to neutral esters; the caloric value of bio-oil increased by 32.26% and 31.64%, respectively; the H2O contents was lowered by 27.74% and 30.87%, respectively; the densities were lowered by 21.77% for both and the viscosities reduced by approximately 97%. After recycling, the resin catalysts showed high catalytic activities and could be used repeatedly. The acid number of bio-oil remarkably decreased by 92.61% after upgrading with 732 resin in the fixed bed reactor. The accelerated ageing test showed improvement of stability, and the aluminum strip corrosion test showed reduced corrosion rate of bio-oil after upgrading. 732 and NKC-9 type ion exchange resins are suitable catalysts for upgrading bio-oils by esterification.

References
[1] Bridgwater A V, Meier D, Radlein D. An overview of fast pyrolysis of biomass. Org Geochem, 1999, 30(12): 14791493. [2] Chang Jie. Research progress in liquefaction technologies of biomass. Modern Chemical Industry, 23(9): 1316. [3] Lu Q, Zhu X F, Li Q X, Guo Q X, Zhu Q S. Biomass fast pyrolysis for liquid fuels. Progress in Chemistry, 19(7/8): 10641070. [4] Bridgwater A V, Peacocke G V C. Fast pyrolysis processes for biomass. Sustainable Renewable Energy Rev, 2000, 4(1): 173. [5] Qin T F. Status and prospects of techniques for making fuel-oil by pyrolysis and chemical liquefaction of biomass. Biomass Chemical Engineering, 2006, 40(suppl): s78s85. [6] Czernik S, Bridgwater A V. Overview of applications of biomass fast pyrolysis oil. Energy Fuels, 2004, 18(2): 590598. [7] Xu J M, Jiang J C, Lu Y J. Progress in research of upgrading of pyrolytic bio-oil. Modern Chemical Industry, 2007, 27(7): 1317. [8] Zhang Q, Chang J, Wang T J, Wu C Z, Xu Y, Zhu X F. Study on

Vous aimerez peut-être aussi