Вы находитесь на странице: 1из 15

44th International Petroleum Conference, Bratislava, Slovak Republic, September 21-22, 2009

PRODUCTION OF DIESEL FUELS FROM WASTE TRIACYLGLYCEROLS BY HYDRODEOXYGENATION J. Mikuleca,*, J. Cvengrob, . Jorkova, M. Bania, A. Kleinovb
b

Slovnaft VRUP, Bratislava, Slovak Republic Faculty of Chemical and Food Technology, Slovak University of Technology, Bratislava, Slovak Republic

Abstract The study is devoted to the issue of direct transformation of triacylglycerols (TAG) to diesel fuels applying a commercially available NiMo and NiW hydrorefining catalysts. It was proved that TAG can be converted to the fuel biocomponent by adding 6.5 % (vol.) of TAG to atmospheric gas oil. In this way, after hydroprocessing at mild conditions (temperature 320360 C, pressure 3.55.5 MPa, LHSV = 1 h-1 and ratio H2:HC = 5001000 Nm3/m3, catalyst presence), gas oil containing 55.5 % of biocomponent was prepared, characterized with standard performance and emission parameters. Long-term stability test of the catalyst was carried out and sufficient catalyst life was confirmed. Performance and emission tests documented that even 5 % (vol.) portion of bio-components reduces the controlled and uncontrolled emissions. Keywords: hydrodesulphurisation, hydrodeoxygenation, decarboxylation, triacylglycerols, atmospheric gas oil, hydrorefining catalysts 1 Introduction Natural triacylglycerols (TAG) present in vegetable oils or animal fats can act as a suitable raw material for producing high-quality engine fuels. Given their high molecular weight and low volatility, they are not appropriate for the use in diesel engines without construction changes of the engines. Transesterification of natural TAG with methanol or ethanol is an industrially applied process for fatty acids methyl and ethyl esters (FAME, FAEE) production. The qualitative parameters of FAME are comparable with fossil diesel quality. FAME are used as an oxygenate components in diesel fuels in up to 5 % vol. The FAME drawback is mainly high price and increased requirements on feedstock quality. The use of FAME is conditioned by a certain adjustment of the equipment when blended with fossil diesel fuel (DF). The fossil fuel blended with FAME is less oxidation-resistant and its long-term storage is not recommended. Direct conversion of TAG to liquid hydrocarbon fuels is a prospective technology of chemical industry. Of the feasible processes such as catalytic cracking and hydrocracking the processes occurring in the presence of hydrogen seem to be more promising. TAG present in vegetable oils and/or animal fats are transformed in the presence of hydrogen and hydrorefining NiMo, CoMo, NiW/-Al2O3 based catalysts are converted to hydrocarbons, mainly to n-alkanes at the temperatures above 300360C and pressure at least 3 MPa leaving propane and CO2 as side-products [1,2,3,5,6]. The mechanism of the reaction is complex and consists of series of consecutive steps, the fastest one being TAG transformation to fatty acids. In the process, three parallel reactions occur: hydrogenation, hydrodeoxygenation and decarboxylation. In hydrogenation-hydrodeoxygenation, n-paraffins with an even number of carbon atoms corresponding to related fatty acids in the used oils/fats, mainly n-C16 and n-C18 are formed along with water and propane. In case of hydrogenation-decarboxylation, the products comprise CO2, propane and n-paraffins with an odd number of carbon atoms in molecules (mainly n-C15 and n-C17), usually lower by one than that in the used TAG acyls [5,6].

44th International Petroleum Conference, Bratislava, Slovak Republic, September 21-22, 2009

Cyclization, aromatization and isomerization are side processes. Increasing the temperature, the decarboxylation rate prevails over that of hydrodeoxygenation [1,20,21]. It has been documented that appropriate selection of the catalytic system allows influencing the main process products [1] as well as the formation of cyclic structures. A degree of unsaturation in TAG chains has an impact on the extent of cyclization in the product. The content of cycloalkanes and alkylbenzenes increases with a degree of unsaturation in the original oil [1]. A higher partial pressure of hydrogen suppresses the formation of olefins, cyclanes and aromates. Since fatty acids are formed in the initial step of TAG conversion, the process was investigated also using model fatty acids and their esters in order to clarify the reaction mechanism. As a catalyst, Pd on various supports was applied [9,10]. Catalytic conversion of the fatty acids has mainly a character of decarboxylation. In case of the fatty acid esters, decarbonylation was the key reaction route [12,13,14]. The product of hydrocracking is usually separated into three fractions by distillation: petrol, diesel fuel, and distillation residue. Due to its high cetane number (CN), the middle diesel fuel distillate is called also Supercetane [17]. Such fuel is comparable with DF, its viscosity is similar to that of FAME. It is miscible with DF in any ratio and is well biodegradable. The high cetane number (CN), 5590, is comparable to that of commercial additives used to increase the CN [23]. An increased CN manifests in emissions reduction (THC, NOx, PM, CO). As its drawback, low-temperature properties caused by a high content of n-alkanes C15C18 should be mentioned. CP (cloud point) and CFPP values range from 20 to 23 C [20]. Catalytic hydroisomerization may be applied to solve the problem. In the function of catalyst, Pt anchored to zeolite HZSM 22 [15] was used. The process occurs at the temperature range of 280370 C, pressure 3.58 MPa, and LHSV 14 h-1. CFPP of the product ranges from -18 to -14 C. Catalyst is, however, sensitive and becomes easily deactivated. The fuel produced through hydrogenation conversion of TAG complies with the existing standards, no new standards are needed. In the process, to obtain a standard-quality product also low-quality oils/fats can be treated [16,18,19] . From the viewpoint of technology, TAG can be processed in an individual unit or to perform its conversion in a blend with light gas oil or vacuum gas oil [19]. An advantage of this mode lies in a lower investment cost [18]. Hydrocracking of vacuum gas oil blended with TAG was successfully carried out, too [20]. The employment of convention refinery technologies as well as hydrorefination catalysts represents also a benefit of hydrogenation cracking [1,2]. In the process, no unusual side products are formed, all of the products are processable in refinery streams. The economy of the process is more favourable than that of transesterification process. It is estimated that the cost of processing (except for that of incoming oil) represents 50 % of transesterification related costs [23]. Investment expenditures of production unit establishment are, however, higher by 50 %. At the time being, the processes of the company Neste Oil [25,26,28] and Ecofinning, jointly offered by UOP and Eni [27,35] are at disposal. The present contribution is aimed at verifying the possibility to transform TAG blended with light gas oil in the process of hydrogenation refining. This mode would offer an advantage of lower investment cost for implementation. Transforming TAG in an individual operation in the absence of crude oil fractions will be studied. By the experiments, available sources of TAG such as rapeseed, sunflower, or palm oils, lard and unconventional source of higher fatty acids such as tall oil will be used. Evaluating the parameters of the obtained products applying common procedures of DF assessment, including performance and emission characteristics and comparing them with those of DF are also the targets of the study.

44th International Petroleum Conference, Bratislava, Slovak Republic, September 21-22, 2009

2 Experimental Catalytic conversion of TAG in a blend (6.5 % vol.) with atmospheric gas oil has been performed in flow apparatus, in a tubular reactor (the total volume 250 ml) with catalytic bed of 100 ml, feed range 100 1 000 ml/h, maximal operation temperature 600 C, maximal operation pressure 100 bar. The device was equipped with the regulation of pressure, temperature, feedstock loading and reaction products discharging. Feedstock container and pipelines were heated to decrease the viscosity and to maintain the possibility of animal fats feeding. Reaction feedstock is pumped by a piston pump and mixed with reaction gas on the head of the reactor. Formed mixture, depending on the amount of the catalyst, passes through a bed, which is, based on the reaction conditions, placed in the reactor body, where the reaction proceeds. The formed product passes subsequently through a water cooler to a separator, where reaction gas is separated from the product. Liquid sample is withdrawn continuously, as this equipment has individual level gauge that enables to set a sample amount. Reaction gas, after being discharged from the separator, passes through a gas flow meter allowing both controlling and measuring its amount. Hydrogen containing gas products were released to the atmosphere by gas meter. 2.1 Feedstock In the first test series, optimal conditions for conversion of TAG (fatty acid composition acyl profile see Tab. 1) to hydrocarbons were searched for. In the second part, the optimal conditions were verified using various feed stocks differing in acyl profiles. In the third part of tests, a long-term test of catalyst stability and production of higher amount of sample for application tests were performed. Among the sources of acyls, the following feed stocks were used: refined rapeseed oil, refined sunflower oil, lard, palm oil and crude tall oil.
Table 1 Acyl profile of materials used by hydroprocessing study, wt. % Fatty acid C14:0 C14:1 C16:0 C16:1 C18:0 C18:1 C18:2 C18:3 C20:0 C20:1 C22:0 C22:1 C24:0 C24:1 Rapeseed oil, refined 0.06 0.00 4.64 0.24 1.96 63.47 20.01 6.97 0.60 1.18 0.15 0.07 0.13 0.14 Sunflower oil, refined 0.07 0.00 6.15 0.07 3.80 22.09 66.62 0.12 0.25 0.23 0.05 0.08 0.03 0.18 Palm oil, refined 1.0 0 35.4 0.3 3.8 45.1 13.4 0.3 0.3 0 0 0 0 0 Lard 1.5 0 31.2 0 16.5 42 6.6 0

The ratio in the first column of Tab. 1 indicates number of carbon atoms:number of double bonds. At the experiments, tall oil with acid value of 139 mg KOH/g was also used. The share of fatty acids was about 41 % wt., the share of rosin acids was about 32 % wt. and neutrals (unsaponifiables, esters and other non-acidic components) of about 27 % wt.

Crude vegetable oils and animal fats contain 9597 % wt. of TAG. Minority components are mono- and diacylglycerols, phospholipides (lecithines 12 %), free fatty acids (0.30.7 %), unsaponifiable matter like tocopherols (0.10.2 %), sterols (about 0.3 %) and terpene hydrocarbons (squalene), then traces of metals (Ca, Mg, Fe, Cu). Refined oils contain more than 99 % wt of TAG; the content of minority components is rapidly lowered (traces of

44th International Petroleum Conference, Bratislava, Slovak Republic, September 21-22, 2009

metals Ca and Mg about 1 ppm, Fe about 0.1 ppm, Cu about 0.03 ppm). Tall oil is obtained as a by-product of the Kraft process of wood pulp manufacture. It contains free fatty acids (mainly oleic and linolic acids) 3050 %, rosin acids like pimaric and abietic acid 4060 % and unsaponifiable matter 1015 % containing sterols (24 %), fatty alcohols, phenols and hydrocarbons. Vacuum distillation in wiped film evaporator is usually used for crude tall oil fractionation to the enriched fatty acid fraction, rosin acid fraction and tall pitch. The common hydroprocessing of atmospheric gas oil produced (see Tab. 2) by crude oil distillation, 5 % (vol.) or 6.5 % (vol.) of refined rapeseed oil was final possibility to prepare renewable diesel. Common hydrorefining and hydrodeoxygenation was carried out at temperature 320360 C, pressure 3.55.5 MPa, LHSV = 1 h-1 and ratio H2:HC = 5001000 Nm3/m3. During the tests procedure, the formed hydrogen sulphide was stripped off by nitrogen. In one series of experiment the feedstock was diluted with an inert solvent isooctane.
Table 2 Atmospheric gas oil properties Characteristic Distillation range Density, 15 C Sulphur content Cetane number Flash point Cloud point CFPP Pour point
o o o o

Unit
o

Value 193371,8 851.9 7613 53.2 74.5 -4 -3 -10

C
3

kg/m

mg/kg C C C C

2.2 Catalysts As a basic benchmark catalyst, the commercial NiMo/-Al2O3 catalyst in sulphidic form was used. The catalyst is applied to desulphurize gas oils used in production of DF containing less than 10 mg/kg of sulphur. Sulphurization was performed directly in a reactor in a stream of hydrogen at a pressure 3 MPa using 5 % solution of dimethyldisulphide in gas oil. The catalyst was dried at 120 C in a stream of nitrogen. The temperature gradually increased with the gradient 100 C/h up to 350 C, it was kept for 1 h at 250 C and for 4 h at 350 C. Sulphurization was successfully carried out in an autoclave, too. Moreover, the catalysts NiMo (6 % NiO, 25 % MoO3) and NiW (6 % NiO, 25 % WO3) were prepared. The catalysts were prepared by impregnation of a support with the solutions of nickel (II) acetate, sodium molybdate (VI) or sodium tungstate (VI). Subsequent to drying at 120 C, the prepared catalysts were calcinated for 4 h at 550 C. Supports TiO2, ZrO2, Al2O3 were purchased from the company Eurosupport Manufacturing Czechia, Litvnov, CZ, NaY was prepared in Slovnaft VRUP, Bratislava. The oxide-based catalysts were converted to the corresponding sulphides by the above mentioned procedure in autoclave. Tab. 3 provides basic characteristics of the catalysts used in testing. The specific surface area was determined with an instrument ASAP2400 (Micromeritics). Activation of samples before measurement: temperature 350 C, vacuum 2 Pa, duration 12 h. Specific surface area SBET was calculated from linearized BET izotherm in a standard range of relative pressure p/p0 (0.050.30). The total volume of pores V0.99 was determined from the adsorbed volume of nitrogen at relative pressure p/p0 = 0.98. Specific volume of micropores Vmicro and specific surface of mesopores (+ external catalyst surface) St was calculated by the t-line method in the range of t (0.350.5 nm). Acidity was determined using the method of temperature-programmed desorption of ammonia (TPDA). Samples (300 mg, 0.10.3 mm) activation before measurement: in a stream of helium up to 500 C.

44th International Petroleum Conference, Bratislava, Slovak Republic, September 21-22, 2009

The catalyst NiW/NaY contains 45 % of micropores and the rest are mezopores from binder. The catalysts on basis TiO2 and ZrO2 are practically pure mezopores ones. Al2O3 catalyst contains only minimal share of micropores.
Table 3 Adsorption properties and acidity of investigated catalysts Catalyst Specific surface area, SBET (mg/g) Specific volume area of micropores, Vmicro (cm3/g) Specific surface area of mesopores, St (cm2/g) Total volume of pores, V0,98 (cm3/g) Pore diameter (A) Acidity TPDA (mmol H /g)
+

NiW/NaY 207 0.092 31.3 0.203 1001000 0.190

NiW/TiO2 59 0.002 54.0 0.287 300 0.120

NiW/ZrO2 159 0.000 159 0.420 85 0.164

NiMo/TiO2 28 0.002 24.9 0.179 240 0.012

NiMo/Al2O3 193 0.005 181 0.486 190 0.320

2.3 Analyses Reaction gas was sampled to bags and analyzed using gas chromatography according to UOP 539-87 with a Shimadzu instrument GC 17 A fy Shimadzu. In the system comprising three columns and two switch valves, oxygen, nitrogen, CO2, CO, hydrogen and light hydrocarbons (methane to n-pentane) are separated. Hydrocarbons heavier than n-pentane elute in one peak. Types of columns: molecular sieve, precolumn: 5m x 0.53mm x 3m SE 54, analytical column: 60 m x 0.53 mm SILICA PLOT. Distribution by boiling points (simulated distillation) was determined according to standard ASTM D 2887 with a AMS 94 device, column: RMX1 15m x 0.53 mm x 2.65 m, detector: FID. GC analyses of liquid products were performed according to standard ASTM D 5134 using a device TRACE GC 2000 INSTRUMENT. Column: WCOT FUSED SILICA 50m x 0.32 mm x 1.2 m CP SIL 5CB, detector: FID. Because the products contained mainly n-alkanes, it was advantageous to use ASTM D 5442 method for evaluation. FIA analysis was used to prove the unsaturated hydrocarbons presence. Evaluation of DF was done using methods prescribed in standard STN EN 590. 2.4 Engine and emission tests Measurements of the performance and emission characteristics were realized using a vehicle VW Touareg R5 2.5 UI (Unit Injection System), year of production 2007. Basic characteristics of the engine are given in Tab. 4.
Table 4 Engine specification of VW Touareg R5 (UI) Number of cylinder Bore (mm) stroke (mm) Volume (L) Compression ratio Maximal load (kW/rev) Maximal torque (N m/rev) 5 81 95.5 2.5 19.5:1 128/3500 400/2000

Measurements of performance parameters were carried out using the chassis dynamometer MAHA LPS 2000 (MBH Haldenwang/Allgu, Germany). Emission measurements were performed with an exhaust gases analyser MAHA MGT5 by means of the emission determination at steady-state regime during idle running and the constant speeds of 60, 90 or 120 km/h. Diesel engine opacity determination was performed by the method of free acceleration with a dynamometer AVL DiSmoke 435. Along with measurements of regulated

44th International Petroleum Conference, Bratislava, Slovak Republic, September 21-22, 2009

emissions, determination of unregulated emissions, namely of VOC (volatile organic compounds) and carbonyls were performed. VOC were determined with an analyser Bernath Atomic equipped with a FID detector. Aldehydes and ketones were determined as described in [36,37]. 3 Results and discussion 3.1 Effect of temperature, pressure and space velocity on catalytic transformation of rapeseed oil The results of rapeseed oil hydrodeoxygenation (HDO) are gathered in Tab. 5 and 6. With increasing temperature at the constant hydrogen pressure, the content of n-alkanes with a number of carbon atoms lower by one as that in the raw material increases, which documents a dominance of reaction (1), i.e. fatty acid hydrodecarboxylation (HDC). A rise in temperature leads also to a higher extent of secondary reactions yielding aromatics, mostly those with one aromatic ring. Comparing the chromatograms obtained using the phases differing in polarity it was shown that olefins were separated and identified. With raising temperature and pressure, the content of olefins decreases. A higher partial pressure of hydrogen at the constant temperature prefers reactions of HDO, suppresses olefin formation in favour of alkanes, the formation of aromatics is not significantly influenced. It exhibits a favourable impact on fatty acids transformation to hydrocarbons. A space velocity decrease exhibits a similar effect.
Table 5 Hydroconversion of rapeseed oil, effect of temperature Hydroconversion conditions: Temperature, C 330 340 Pressure, MPa 3 -1 LHSV, h 1 H2:TAG, Nm3/m3h 250 Product: < n-C14 n-C15 n-C16 n-C17 n-C18 > n-C18 Isoalkanes, cycloalkanes, olefins Aromatics Polar substances 350

Composition, % wt.: 0.26 0.31 0.35 2.84 3.09 3.35 2.51 2.43 2.27 49.35 51.16 53.32 35.95 33.40 29.09 1.51 1.54 1.50 5.59 0.95 1.06 6.34 1.27 0.45 8.31 1.54 0.26

At the temperature above 360 C, space velocity 0.81 h-1, pressure of hydrogen at least 4.5 MPa and the ratio H2:TAG = 1000 and more, the TAG conversion is quantitative and does not lead to polar substances and olefins. Using other catalysts, the above conditions may be milder. We have also tried out to apply hydrogen in high excess and to use an inert solvent isooctane. Such a change has a positive effect on the course of HDC and HDO reactions. As can be seen in Fig. 1, the dependence of HDC and HDO reactions on temperature and pressure is similar but HDO reactions are more favored under the same conditions (temperature, pressure).

44th International Petroleum Conference, Bratislava, Slovak Republic, September 21-22, 2009

Table 6 Hydroconversion of rapeseed oil, effect of pressure Hydroconversion conditions: Pressure, MPa 3.5 4.0 Temperature, C 340 LHSV, h-1 1 H2:TAG, Nm3/m3h 250 Product: < n-C14 n-C15 n-C16 n-C17 n-C18 > n-C18 Isoalkanes, cycloalkanes, olefins Aromatics Polar substances 4.50

Composition, % wt.: 0.40 0.27 0.37 3.81 2.54 2.86 3.39 2.70 3.00 49.12 49.28 46.26 34.06 37.12 38.94 0.84 1.80 1.29 7.02 1.15 0.20 5.03 1.09 0.18 5.96 1.17 0.13

2,0
ratio n-C17/n-C18, p=4,5 MPa, H2:TAG=1000, diluted with isooctane

1,8 1,6 1,4 ratio n-C17/n-C18 1,2 1,0 0,8 0,6 0,4 0,2 290

ratio n-C17/n-C18, p=5,5 MPa, H2:TAG=1000, diluted with isooctane ratio n-C17/n-C18, p=3 MPa, H2:TAG=250

300

310

320

330

340

350

360

370

380

390

reaction temperature, oC

Figure 1 Hydroconversion of rapeseed oil. Temperature dependence of n-C17/n-C18 ratio

Mechanism of TAG conversion over hydrotreating catalyst in the presence of hydrogen and elimination of oxygen from a TAG molecule is an intricate process that needs to be verified using a simpler system. Real raw materials contain also other compounds that may have an effect on the catalytic system. In the initial stage of our experiments we were evaluating, in a continuous reactor with NiMO/Al2O3 as the catalyst, the effect of temperature, hydrogen pressure and space velocity on the course of chemical reactions. We worked intentionally also in the range below the process optimum parameters. Rapeseed oil

44th International Petroleum Conference, Bratislava, Slovak Republic, September 21-22, 2009

was used as a model substance for TAG. It has a high content of acyls, which possess unsaturated bonds. Basic reactions leading to required products formation can be written down as follows [1]: Decarboxylation: CnH2n+1COOH CnH2n+2 + CO2 Decarbonylation: CnH2n+1COOH + H2 CnH2n+2 + H2O + CO Reduction: CnH2n+1COOH + 3H2 Cn+1H2n+4 + 2H2O (3) Reaction products are gaseous hydrocarbons, mixture of liquid hydrocarbons with higher n-alkanes predominating (C15-C18), organic compounds (in part solid) as intermediary products of catalytic reactions (saturated TAG, fatty acids, alcohols, fatty acid esters) and reaction water. The gaseous phase contains CO2, CO, propane, methane, ethane and propylene. Their mutual proportion depends on reaction conditions, raw material used and partly also on the catalyst type. Fig. 2 shows the possible reaction mechanism [1]. Hydrogenation of double bonds in acyls occurs at temperatures lower than those of HDC and HDO processes; reaction heat is released depending on the raw material composition. The gas sample composition indicates the reaction course. In the first step, water is eliminated and monocarboxylic acids are formed without degradation of the long alkane chain. The next step involves HDC and/or HDO of the fatty acid accompanied by the formation of hydrocarbons. It is obvious that the former reaction runs faster since along hydrocarbons free monocarboxylic acids C16 and C18 were identified in the liquid phase. (2) (1)

Figure 2 Flow-chart of processes of TAG conversion in the presence of hydrogen and NiMo/Al2O3 catalyst

The HDC reaction is favored by lower partial pressures of hydrogen and by higher temperatures. HDC involves CO2 elimination and formation of n-alkane with an odd number of carbon atoms (C17). Increasing the hydrogen partial pressure shifts reactions towards reduction and HDC and/or HDO with formation of n-alkanes having an even number of carbon atoms (C18), and propane, water and CO as byproducts. In addition, methane and ethane were observed in low concentrations. Methane is formed by a side equilibrium reaction of hydrogen and CO. Cracking of alkanes thus formed occurs at higher temperatures, exceeding 380 C, and when stronger acidic catalytic centers are present.

44th International Petroleum Conference, Bratislava, Slovak Republic, September 21-22, 2009

The ratio of n-alkanes with an odd number of carbon atoms to n-alkanes with an even number of carbon atoms can serve as an indicator for assessing the reaction course of HDC and HDO. This ratio has a marked effect on hydrogen consumption, and hence on the reactors caloric balance as well as on CO2, CH4, and CO content in the hydrogen gas; it also affects the activity of the catalyst. In addition to principal reactions of HDC and HDO, competitive reactions of isomerization and alkylation of byproducts occur, too. 3.2 Effect of the feedstock type on catalytic hydroprocessing of TAG In Tab. 7, the results of comparison of four different feedstocks with different TAG composition are presented. Stemming from the results it is obvious that the feedstock composition has a significant effect on n-alkanes distribution in the product. The portion of iso-alkanes was negligible when using the investigated catalyst. Aromates concentration is similar in all studied products and it lies below 2 % wt. level.
Table 7 Hydroconversion of various feedstocks Type of TAG Temperature, C Pressure, MPa LHSV, h-1 H2:TAG Product < n-C14 < iso-C14 n-C15 iso-C15 n-C16 iso-C16 n-C17 iso-C17 n-C18 iso-C18 > n-C18 > iso-C18 Aromatics Polar substances
o

Rapeseed oil 340 4.5 0.8 500 0.46 0.03 2.00 0.02 2.97 0.04 35.26 1.60 47.98 2.54 3.44 1.89 1.77 0.00

Palm oil

Sunflower oil

Lard 350 4.5 0.8 250 1.4 5.65 16.22 16.57 4* 43.67 2.78 n.a. 9.65

Hydroconversion conditions: 350 350 4.5 4.5 0.6 0.8 250 500 Composition, % wt.: 2.62 3.71 0.33 0.27 15.46 3.40 0.09 0.01 20.86 3.40 0.09 0.04 23.97 35.34 0.71 1.52 30.40 41.91 1.11 0.97 1.52 1.55 1.08 0.64 1.76 1.86 0 5.39

*sum of all isoalkanes , n.a. not analyzed

3.3 Hydroconversion of TAG with atmospheric gas oil and effect of type of the catalyst on catalytic transformation of TAG Interesting results were achieved by testing the NiW/NaY catalyst with RO (Tab. 8). The catalyst was exceptionally active even at relatively low temperatures and low pressure. In the parameter range analyzed, HDC was predominant; at 340 C a considerable quantity of olefins C15-C18 was formed and saturated fatty acids were present among reaction products. The gaseous phase contained substantial portions of CO and propylene. It is obvious that catalyst supports containing a mixture of micro- and macropores can significantly affect reactions of hydrogenation transformation of TAG into hydrocarbons.

44th International Petroleum Conference, Bratislava, Slovak Republic, September 21-22, 2009

10

Table 8 Hydroconversion of rapeseed oil on NiW/NaY catalyst Temperature, C Pressure, MPa LHSV, h-1 H2:TAG, Nm3/m3h Product: n-C15/n-C16 n-C17/n-C18 n-C17/(n-C17+n-C18) Aromatics Polar substances 340 360 3.5 1 1000 Composition, % wt.: 3.22 3.68 298 4.12 4.16 4.81 0.81 0.83 0.83 1.30 5.30 1.04 5.02 0.54 0.31 320

HDS of petroleum fractions is a common technological operation in an oil refinery. We have verified experimentally the option of common desulphurization of distillate light gas oil in the mixture using a selected TAG. We were working with relatively low TAG contents in the mixture to have approximately 5 % of biocomponents in the product. 6.5 % (V/V) of the selected TAG was mixed with unprocessed gas oil from petroleum distillation. The process was tested using NiMo and NiW catalysts on various catalyst supports. The results are presented in Tab. 9. Over all catalysts tested all the reactions occurred in parallel; what varied were only mutual proportions of the reactions depending on the respective technological conditions and catalyst support properties.
Table 9 Hydroconversion of the blends AGO with RO, TO and T (6.5 % vol.) over different catalysts
catalyst temperature, oC pressure, MPa LHSV, h-1 H2:HC+TAG NiMo/ Al2O3 380 5.5 1 1000 NiMo/ Al2O3 380 5.5 1 1000 NiMo/ Al2O3 380 5.5 1 1000 NiMo/ TiO2 360 3.5 1 1000 NiW/ NaY 360 3.5 1 1000 AGO+RO 3.80 -3 55.7 4.3 25.8 NiW/ TiO2 360 3.5 1 1000 AGO+RO 2.40 +2 55.2 4.9 26.5 NiW/ ZrO2 360 3.5 1 1000 AGO+RO 0.94 +4 56.7 5.2 24.9

AG feed AGO+RO AGO+TO AGO+T AGO+RO O 1.64 0.70 0.84 2.32 n-C17/n-C18 -3 -4 +5 -2 +2 CFPP, oC Cetane number 53 57 56.4 56.8 55.4 PAH, % wt. 6.9 3.4 4.1 3.5 4.8 aromatics, % 27. 23.9 26.3 23.8 26.2 wt. 7 AGO- atmospheric gas oil, RO- rapeseed oil, TO crude tall oil, T- tallow

It was proved that TAG conversion in the presence of unprocessed gas oil has certain characteristic features: - TAG conversion to hydrocarbons (reactions 1 and 2) consumes a large amount of hydrogen, which in turn generates heat, - HDO reactions are fast, competing with sulfur elimination reactions from sulfur-containing compounds, - the process must take place with hydrogen in high excess to eliminate the formation of insoluble waxes, coke and polymeric deposits on the catalyst, - increasing temperature promotes the HDC reaction, giving rise to CO2; an important side reaction is a reversible reaction of hydrogen with carbon dioxide giving rise to carbon monoxide: CO2 + H2 CO + H2O, - gas products also include methane formed by methanization from CO and hydrogen on the metallic component of the catalyst, - CO2 and CO being formed have an inhibition effect on the HDS process, - gases formed in the reaction reduce the partial pressure of hydrogen, which is a limiting factor especially when hydrogen pressure is low.

44th International Petroleum Conference, Bratislava, Slovak Republic, September 21-22, 2009

11

The blend of AGO+TO shows low value of the ratio C17/C18 (0.70). That means the decarboxylation occurs here in lower extent compared to that of RO. This can be caused by lower portion of unsaturated acyls of fatty acids in TO. Such conclusion is confirmed also by similar low ratio C17/C18 in the case of AGO+T (0.80). Rosin acids present in TO can show different behavior from linear fatty acids during the hydroconversion. HDO/HDC at temperatures higher than 360 C, LHSV 0,81 h-1, hydrogen pressure at least 4.5 MPa and the ratio H2:TAG = 1000 and more was fully completed over all catalysts. Catalyst acidity had no significant effect on isomerization of n-alkanes being formed. Differences in CP and CFPP values were small. As the reaction gave rise to higher n-alkanes with a higher melting point, it was helpful that gas oil worked as a solvent. Ni W-containing catalysts exhibit an extraordinary hydrogenation activity and at lower reaction temperature (320340 C) wax like insoluble product were formed. In cases of catalyst supports with low specific surface (ZrO2), the level of desulphurization was much lower when compared to the commercial catalyst. At temperatures 360380 C and with high hydrogen excess, a greater quantity of reaction water was formed, pointing out to a more significant proportion of the HDO reaction. 3.4 The quality of products, performance and emission tests of diesel fuel with renewable components from hydroprocessing Hydroconversion of the blend AGO and TAG was proved by long-term experiment. Catalytic activity did not significantly change after two weeks of test duration. HDS and HDC reaction are competitive and so the desired desulphurization was not achieved. All other parameters met the standard EN 590. Because of hydrogen excess in the reaction the hydrodearomatization of aromates and polyaromates occured. The result product showed increased cetane number, which has positive effect on emissions, aromate content and density. This fact is very important for the refinery due to possibility to blend low-value stream to diesel. Biocomponent content has no negative influence on de-emulsification properties of DF and its corrosiveness. Oxidation stability was slightly lowered but the sample did not contain antioxidants. Performance characteristics of blended fuel DF and 5 % HDO from RO were comparable to those of neat DF. This is the result of presence similar components in both compared fuels. The results of emission characteristics of the DF containing 5 % HDO from RO in comparison to DF are presented in Tab. 10. An addition of the second generation biocomponent to DF exhibits a significantly positive effect on all monitored emissions, those unregulated in particular.
Table 10 Emission characteristics of tested fuels Fuel Testing conditions idling 60 km.h-1 90 km.h-1 120 km.h-1 idling 60 km.h-1 90 km.h-1 120 km.h-1 VOC mg/kg 4.5 6.0 4.0 14.0 2.4 2.5 1.9 2.0 Corg. mg.m-3 7.2 9.7 6.4 22.5 3.9 4.0 3.1 3.2 NOx mg/kg 44.0 87.6 294.6 476.2 28.2 85.6 288.0 474.6 Carbonyls g/l 0.9 2.8 4.8 0.2 0.2 0.8 0.8 0.8

DF

95 % DF + 5 % HDO from RO

DF diesel fuel, HDO from RO hydrodeoxygenate from rapeseed oil

44th International Petroleum Conference, Bratislava, Slovak Republic, September 21-22, 2009

12

4 Proposal of new technology for production diesel fuel with renewable components A new concept of co-processing AGO and TAG will allow producing the diesel fuel with renewable diesel share. Technology allows using non edible feedstock in the production diesel fuel that can meet EN 590 specification. Another advantage of this process is fulfillment of the governmental mandate on using bio-components in diesel pool. The standard refinery distillate hydrotreating units do not appear to be suitable for renewable diesel production in a co-processing scheme, but co-processing of biomass such as vegetable oils and/or animal fats may be processed in a revamp refinery distillate hydrotreating unit. Flow scheme is in Fig. 3, the proposal of common HDS of the AGO and TAG is in Fig. 4. The advantage of the technology is the opportunity to use cheaper feedstock like TAG wastes. Another advantage of this process is fulfillment of the governmental mandate on using bio-components in diesel pool. But technology must be carefully designed and the problems accompanied with materials, catalysts and heat of reaction must be solved. Vegetable oils/animal fats contain free fatty acids and the impact of their acidity to the reactor and pipes corrosion should be taken into account. Moreover, they contain traces of metals and phosphorus acting as catalytic poisons cutting down the catalyst activity and lifetime. The water formed during HDO poses a further problem to the catalyst. Therefore, it is advantageous to perform the reaction in a separate reactor connected in line with the main hydrorefining reactor. The issue of the catalyst choking up with metals and phosphorus can be solved using a catalytic bed filled with a cheaper catalyst trapping the mentioned adverse substances prior to the main catalyst. The concept as designed will enable the use of the refinerys infrastructure. A key factor to success is selecting an appropriate combination of catalysts to ensure gas oil HDS at the requested sulphur level below 10 mg/kg. The catalyst must be powerful enough to allow transforming TAG to hydrocarbons. Because of a need for high hydrogen excess and great formation of heat, it is necessary to resolve heat transfer in the process. Also important is removal of CO2, CO, and CH4 from the hydrogen recycle. CO2 can be removed by amine washing or by using the PSA system, if installed in the refinery. CO can be eliminated in the PSA system separately, or following catalytic methanization. The optimum solution will be based on the specific situation prevailing in the refinery. The technology represents green technology which is able to produce diesel fuel of second generation from non-edible sources.
Figure 3 Renewable diesel production co-processing of biofeed in hydrodesulphurization unit
Hydrogen

Atmospheric gas oil

- hydrodesulphurization of gas oil ULSD

Hydrogen

- pretreatment, cleaning Oil/fat

- hydrotreatment - hydrodeoxygenation and decarboxylation - hydrodesulphurization - hydrodewaxing

Solids

AGO

44th International Petroleum Conference, Bratislava, Slovak Republic, September 21-22, 2009
VEG OIL/FATS

13

CATALYST PSA

RECYCLE HYDROGEN COMPRESSOR CO2

HEATER REACTOR

AMINE ABSORBER

LIGHT ENDS

SEPARATOR

STRIPPER

H2 SOUR WATER HYDROGEN MAKE-UP COMPRESSOR RENEWABLE DIESEL

Figure 4 Flow scheme proposal for co-processing hydrodesulphurization of atmospheric gas oil with hydrodeoxygetion and hydrodecaroxylation of renewable biocomponent

5 Conclusions TAG can be converted to the second generation biocomponents in mixtures with atmospheric gas oil from crude oil distillation using a hydrorefining catalyst. Adding 6.5 % (vol.) of vegetable oil, gas oil containing 55.5 % of biocomponent was prepared, characterized with performance and emission parameters similar to fossil diesel. The philosophy of introduction of a second reactor into the existing structure of hydrorefining unit allows reducing investment cost of biocomponent processing. The desulphurization of atmospheric gas oil is slower than n-alkane production from TAG over typical hydrodesulphurization catalyst. Key factor of successful hydrodesulphurization and hydrodeoxygenation/hydrodecarboxylation is in appropriate selection of the catalyst and technological condition. The selectivity to hydrodeoxygenation/hydrodecarboxylation products increases with increasing temperature. Acknowledgement The authors thank Pavol Kuna (Slovnaft VRUP, Bratislava) for performing catalytic tests, assoc. prof. Pavol Hudec, PhD. (STU, Bratislava) for performing catalyst parameters. This work was supported by the Slovak Research and Development Agency under the contract No. APVV-20-037105. Nomenclature: AGO atmospheric gas oil, CFPP cold filter plugging point, CP cloud point, DF diesel fuel, HDC hydrodecarboxylation, HDO hydrodeoxygenation, HDS hydrodesuplhurization, LHSV liquid hourly space velocity, PAH polyaromatic hydrocarbons, PSA pressure swing adsoprtion, RO rapeseed oil, T tallow, TAG triacylglycerols, TO tall oil, ULSD ultra low sulphur diesel, VOC volatile organic compounds References [1] [2] da Rocha Filho DN, Brodzki D, Djga-Mariadassou G. Formation of alkanes, alkylcykloalkanes and alkylbenzenes during the catalytic hydrocracking of vegetable oils. Fuel 1993;72(4):543549. Gusmao J, Brodzki D, Djga-Mariadassou G, Frety R. Utilization of vegetable oils as

44th International Petroleum Conference, Bratislava, Slovak Republic, September 21-22, 2009

14

[3] [4] [5] [6] [7] [8] [9] [10] [11] [12] [13] [14] [15]

[16] [17] [18] [19] [20] [21]

an alternative source for diesel-type fuel: Hydrocracking on reduced Ni/SiO2 and sulphided Ni-Mo/-Al2O3. Catalysis Today 1989;5(4):533544. da Rocha Filho D N, Bentes M H S, Brodzki D, Djga-Mariadassou G. Catalytic conversion of Hevea brasiliensis and Virola sebifera oils to hydrocarbon fuels. Journal of the American Oil Chemists' Society. 1992;69(3):266271. Nunes PP, Brodzki D, Bugli G, Djega-Mariadassou G. Hydrocraquage sous pression d'une huile des soja: Procd d'tude et allure gnrale de la transformation. Revue de l'Institute Franais du Ptrole. 1986;41(3):421430. imek P, Kubika K , ebor G , Pospil, M. Hydroprocessed rapeseed oil as a source of hydrocarbon-based biodiesel. Fuel 2009;88(3):456460. Kubika D, imek P, ilkova N. Transformation of Vegetable Oils into Hydrocarbons over Mesoporous-Alumina-Supported CoMo Catalysts. Topics in Catalysis 2009;52(1-2):161168. Kubikov I, Snre M, Ernen K, Mki-Arvela P, Murzin DYu. Hydrocarbons for diesel fuel via decarboxylation of vegetable oils. Catalysis Today 2005;106(1-4):197200. Lestari S, Mki-Arvela P, imakova I, Beltramini J, Max Lu GQ, Murzin DY. Catalytic Deoxygenation of Stearic Acid and Palmitic Acid in Semibatch Mode. Catalysis Letters 2009;130(1-2): 4851. Pivi Mki-Arvela, Kubikov I, Snre M, Ernen K, Murzin DY. Catalytic deoxygenation of fatty acids and their derivatives. Energy and Fuel 2007;21(1): 3041. Lestari S, Mki-Arvela P, imakova I, Beltramini J,Max Lu GQ, Murzin DYu. Synthesis of Biodiesel via Deoxygenation of Stearic Acid over Supported Pd/C Catalyst. Catalysis Letters 2008;122(3-4):247251. Do PD, Chiappero M, Lobban LL, Resasco DE. Catalytic Deoxygenation of MethylOctanoate and Methyl-Stearate on Pt/Al2O3. Catalysis Letters 2009; 130(1-2):918. enol O, Ryymin EM, Viljava T-R, A.O.I. Krause AOI. Reactions of methyl heptanoate hydrodeoxygenation on sulphided catalysts. Journal of Molecular Catalysis A: Chemical 2007;268(1-2):18. enol O, Viljava T-R, A.O.I. Krause AOI. Effect of sulphiding agents on the hydrodeoxygenation of aliphatic esters on sulphided catalysts. Applied Catalysis A: General 2007;326(2):236244. enol O, Viljava T-R, A.O.I. Krause AOI. Hydrodeoxygenation of aliphatic esters on sulphided NiMo/-Al2O3 and CoMo/-Al2O3 catalyst: The effect of water. Catalysis Today 2005;106(1-4):186189. Hancsk J, Krr M, Magyar Sz, Boda L, Holl A, Kall D. Investigation of the production of high cetane number bio gas oil from pre-hydrogenated vegetable oils over Pt/HZSM-22/Al2O3. Microporous and Mesoporous Materials. 2007;101(12):148152. Huber GW, O'Connor P, Corma A. Processing biomass in conventional oil refimeries:Production of high quality diesel by hydrotreating vegetable oils in heavy vacuum oil mixtures. Applied Catalysis A: General 2007;329:120129. Stumborg M, Wong A, Hogan E. Hydroprocessed vegetable oils for diesel fuel improvement. Bioresource Technology 1996;56(1):1318. Donnis B, Egeberg RG, Blom P, Knudsen KG. Hydroprocessing of Bio-Oils and Oxygenates to Hydrocarbons.Understanding the Reaction Routes. Topics in Catalysis (2009);52(3):229240. Jerzy Walendziewski J, Stolarski M, uny R, Klimek B. Hydroprocesssing of light gas oil rape oil mixtures. Fuel Processing Technology 2009;90(5):686691. Bezergianni S, Kalogianni A, Vasalos IA. Hydrocracking of vacuum gas oil-vegetable oil mixtures for biofuels production. Bioresource Technology. 2009;100(12):30363042. Furimsky E. Chemistry of Catalytic Hydrodeoxygenation. Catal. Rev.Sci Eng. 1983; 25(3):421458.

44th International Petroleum Conference, Bratislava, Slovak Republic, September 21-22, 2009

15

[22] [23]

[24] [25] [26] [27] [28] [29] [30] [31] [32] [33] [34] [35] [36] [37]

Furimsky E. Catalytic hydrodeoxygenation. Applied Catalysis A: General 2000;199(2):147190. Soveran D, Sulatisky M, Ha K, Robinson W, Stumborg MA. The effect on diesel engine emissions with high cetane additives from biomass oils. Proc. American Chemical Society (Division of Fuel Chemistry) Meeting, San Francisco, CA, 5-10 April 1992. Craig WK, Soveran DW. U.S. Patent 4,992,605. Production of hydrocarbons with a relatively high cetane rating , February 12, 2001. European patent 1 396 531(Fortum OYJ). Process for producing a hydrocarbon component of biological origin. Patent WO 2004/022674 (Fortum OYJ). Diesel fuel composition, comprising components based on biological raw material obtained by hydrogenating and decomposing fatty acids. European patent 1 728 844 (UOP). Production of diesel fuel from biorenewable feedstocks. Rantanen L, Linnaih R, Aakto P, Harju P. NExBTL-Biodiesel fuel of the second generation, SAE-2005-01-3771. Michaelsen NH, Egeberg R,Nystrm S. Consider New Technology to Produce Renewable Diesel, Hydrocarbon Processing 2009;88(2):4142,43. Market T. et al. Oportunities for Biorenewables in Oil refineries, UOP, DOE Award No.:DE-FG36-05GO15085, 2005. US. Patent application 2007/0287871A1. Silicoaluminophosphate Isomerization catalyst , Brevoord et al. , Albermarle Netherlands BV. US. Patent application 2008/066374A1. Reaction System for Producing f diesel fuel from Vegetable and Animal Oils, Herstkowitz M. et al., Buchanan, Ingersol & Rooney PC. US. Patent application 2006/0207166A1. Production of diesel fuel from Vegetable and Animal Oils, Herstkowitz M. et al., Buchanan, Ingersol & Rooney PC. Iki H, Iguchi Y, Koyama A. Applicability if Hydrogenated Palm Oil for Automotive Fuels, 16th Saudi Arabia-Japan Joint Symposium, Dahran, Saudi Arabia, November 56, 2006. Tom N Kalnes T, Marker T, Shonnard DR, Ken P Koers KP. Green diesel production by hydrorefining renewable feedstocks. Biofuels Technology, 2008; Q4, 7-11. Potter W, Karst U. Identification of Chemical Interferences in Aldehyde and Ketone Determination Using Dual-Wavelength Detection. Ana Chem 1996; 68:3354-3358 EPA Air esources Board SOP MLD 104. Standard operating procedure for the determination of aldehyde and ketone compounds in automotive source samples by HPLC.

Вам также может понравиться