single stage hydroprocessing of pyrolysis oil in a continuous packed-bed reactor

6
Single Stage Hydroprocessing of Pyrolysis Oil in a Continuous Packed-Bed Reactor Divya R. Parapati, Vamshi K. Guda, Venkata K. Penmetsa, Philip H. Steele, and Sathish K. Tanneru Department of Sustainable Bioproducts, Mississippi State University, Starkville, MS 39759; [email protected] (for correspondence) Published online 13 February 2014 in Wiley Online Library (wileyonlinelibrary.com). DOI 10.1002/ep.11954 Raw bio-oil cannot be combusted as transportation fuel directly because of its high acidity, high water content, lower heating value, and variable viscosity over time. Therefore, bio-oil should be chemically converted to a more stable liquid product before subjecting it to hydrodeoxygenation (HDO) conditions. This research article focuses on catalytic hydro- processing of pretreated bio-oil (PTBO) in a single stage reac- tion using various catalyst compositions in a packed-bed reactor. Four catalysts, a conventional hydrotreating catalyst (CoMo/g-Al 2 O 3 ), an Fe-Cr based mixed oxide catalyst, an FeW/Si-Al catalyst, and a 1:2 mixture of Ru/g-Al 2 O 3 and Ni/Si-Al catalyst, were tested for conversion of the PTBO to mixed liquid hydrocarbons at 350–400 C, 1500 psig hydro- gen pressure, and at a liquid hourly space velocity (LHSV) of 0.2–0.3 h 21 . Liquid products produced from the HDO treat- ments were analyzed for properties such as acid value, heat- ing value, elemental analysis, water content, and chemical characterization. The conventional hydrotreating catalyst, CoMo/g-Al 2 O 3 , performed the best among the four catalysts employed to reduced the acid value to 2 mg KOH/g and oxy- gen content to 0.1% while improving the heating value to 43 MJ/kg of the liquid product. The detailed hydrocarbon analy- sis of the reduced CoMo/g-Al 2 O 3 upgraded hydrocarbon mix- ture showed the presence of olefins, iso-paraffins, followed by naphthenes and aromatics. Simulated distillation results indicated that the liquid fuel had a boiling point range of 69–304 C, indicating the presence of petroleum equivalents of 50% gasoline (38–170 C), 30% jet fuel (170–250 C), and 20% diesel (250–304 C) range hydrocarbons. V C 2014 American Institute of Chemical Engineers Environ Prog, 33: 726–731, 2014 Keywords: bio-oil, pretreatment, upgrading, hydrodeoxy- genation, hydrotreating INTRODUCTION Biomass, due to its high carbon value, renewability, and environmentally benign nature, has attracted interest as a potential alternative fuel resource. Fast pyrolysis, a thermo- chemical technology performed at temperatures from 400 to 550 C in the absence of oxygen, is a potential route to con- vert lignocellulosic biomass to a liquid fuel called bio-oil [1– 4]. The yields of bio-oil obtained from fast pyrolysis range from 60 to 75 dry wt% of wood, depending on process type and conditions [3, 5]. Bio-oil, a dark brown viscous liquid, possesses a high oxygen content in the form of water and a complex mixture of numerous oxygenated chemical func- tionalities including carbonyl groups, acids, alcohols, alde- hydes, esters, ketones, sugars, phenols, phenol derivatives, and a large proportion (20–30 wt%) of lignin derived oligom- ers [6]. Moreover, it possesses relatively low heating value, low volatility, thermal instability, corrosiveness, immiscibility with fossil fuels, and a tendency to polymerize over time, largely due to the presence of a high percentage of these reactive oxygenates [7]. The acids present in the bio-oil pro- mote aldol reactions and also accelerate bio-oil aging [8]. Raw bio-oils have been tested in diesel engines, turbines, and Stirling engines. However, the results have been disap- pointing with objectionable engine erosion, deposits, and sig- nificant wear except in Stirling engines. It is universally agreed that bio-oils must be substantially upgraded/deoxygenated prior to their utilization as engine fuels [5, 9, 10]. Presently, the widely employed bio-oil upgrading methods include hydro- deoxygenation [9, 11–13], catalytic cracking, super-critical treat- ment, and steam reforming [14]. Catalytic cracking can only partially deoxygenate the raw bio-oil and produces low liquid yields [14]. Steam reforming produces a low energy density gaseous fuel and supercritical treatment requires high capital cost due to the requirement of high pressure vessels [15]. Other upgrading methods such as olefination and esterification [16, 17] are used to upgrade bio-oil to boiler fuel. HDO is one of the upgrading methods reported to pro- duce a highly de-oxygenated (as low as zero wt % oxygen in the upgraded product), high energy liquid fuel. Previous HDO studies [4, 18, 19] over nearly three decades have pro- vided considerable information about methods to upgrade bio-oil by this technology. However, rapid catalyst deactiva- tion (by coking), reactor plugging, and low product yields continue to be problematic and further research is required to refine current HDO methods and catalysts. Our studies employed a pretreated bio-oil (PTBO) treated by HDO method to upgrade bio-oil to mixed liquid hydrocarbons. Moreover, instead of employing a two-step upgrading method (hydrotreatment followed by hydrocracking), a single-stage hydroprocessing was applied to produce liquid fuel from bio-oil. It has become customary to practice the hydroprocessing of bio-oil by utilizing a 2-stage approach in which the 1st stage comprises a hydrotreating stage utilizing a mild temper- ature (<300 C) for the reaction. This 1st stage eliminates the polymerization of bio-oil that occurs when raw bio-oil is subjected to high temperatures. Hydrocracking the hydro- treated product is then practiced in a 2nd stage reaction at higher temperature (>350 C). The 2-stage hydroprocessing method usually requires two reactors which increases Presented at the Tcbiomass2013 conference; September 3–6, 2013, at Chicago, Illinois. V C 2014 American Institute of Chemical Engineers Environmental Progress & Sustainable Energy (Vol.33, No.3) DOI 10.1002/ep 726 October 2014

Upload: sathish-k

Post on 07-Apr-2017

212 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Single stage hydroprocessing of pyrolysis oil in a continuous packed-bed reactor

Single Stage Hydroprocessing of Pyrolysis Oil

in a Continuous Packed-Bed ReactorDivya R. Parapati, Vamshi K. Guda, Venkata K. Penmetsa, Philip H. Steele, and Sathish K. TanneruDepartment of Sustainable Bioproducts, Mississippi State University, Starkville, MS 39759; [email protected](for correspondence)

Published online 13 February 2014 in Wiley Online Library (wileyonlinelibrary.com). DOI 10.1002/ep.11954

Raw bio-oil cannot be combusted as transportation fueldirectly because of its high acidity, high water content, lowerheating value, and variable viscosity over time. Therefore,bio-oil should be chemically converted to a more stable liquidproduct before subjecting it to hydrodeoxygenation (HDO)conditions. This research article focuses on catalytic hydro-processing of pretreated bio-oil (PTBO) in a single stage reac-tion using various catalyst compositions in a packed-bedreactor. Four catalysts, a conventional hydrotreating catalyst(CoMo/g-Al2O3), an Fe-Cr based mixed oxide catalyst,an FeW/Si-Al catalyst, and a 1:2 mixture of Ru/g-Al2O3 andNi/Si-Al catalyst, were tested for conversion of the PTBO tomixed liquid hydrocarbons at 350–400�C, 1500 psig hydro-gen pressure, and at a liquid hourly space velocity (LHSV) of0.2–0.3 h21. Liquid products produced from the HDO treat-ments were analyzed for properties such as acid value, heat-ing value, elemental analysis, water content, and chemicalcharacterization. The conventional hydrotreating catalyst,CoMo/g-Al2O3, performed the best among the four catalystsemployed to reduced the acid value to 2 mg KOH/g and oxy-gen content to 0.1% while improving the heating value to 43MJ/kg of the liquid product. The detailed hydrocarbon analy-sis of the reduced CoMo/g-Al2O3 upgraded hydrocarbon mix-ture showed the presence of olefins, iso-paraffins, followed bynaphthenes and aromatics. Simulated distillation resultsindicated that the liquid fuel had a boiling point range of69–304�C, indicating the presence of petroleum equivalentsof 50% gasoline (38–170�C), 30% jet fuel (170–250�C), and20% diesel (250–304�C) range hydrocarbons. VC 2014 American

Institute of Chemical Engineers Environ Prog, 33: 726–731, 2014

Keywords: bio-oil, pretreatment, upgrading, hydrodeoxy-genation, hydrotreating

INTRODUCTION

Biomass, due to its high carbon value, renewability, andenvironmentally benign nature, has attracted interest as apotential alternative fuel resource. Fast pyrolysis, a thermo-chemical technology performed at temperatures from 400 to550�C in the absence of oxygen, is a potential route to con-vert lignocellulosic biomass to a liquid fuel called bio-oil [1–4]. The yields of bio-oil obtained from fast pyrolysis rangefrom 60 to 75 dry wt% of wood, depending on process typeand conditions [3, 5]. Bio-oil, a dark brown viscous liquid,possesses a high oxygen content in the form of water and a

complex mixture of numerous oxygenated chemical func-tionalities including carbonyl groups, acids, alcohols, alde-hydes, esters, ketones, sugars, phenols, phenol derivatives,and a large proportion (20–30 wt%) of lignin derived oligom-ers [6]. Moreover, it possesses relatively low heating value,low volatility, thermal instability, corrosiveness, immiscibilitywith fossil fuels, and a tendency to polymerize over time,largely due to the presence of a high percentage of thesereactive oxygenates [7]. The acids present in the bio-oil pro-mote aldol reactions and also accelerate bio-oil aging [8].

Raw bio-oils have been tested in diesel engines, turbines,and Stirling engines. However, the results have been disap-pointing with objectionable engine erosion, deposits, and sig-nificant wear except in Stirling engines. It is universally agreedthat bio-oils must be substantially upgraded/deoxygenatedprior to their utilization as engine fuels [5, 9, 10]. Presently, thewidely employed bio-oil upgrading methods include hydro-deoxygenation [9, 11–13], catalytic cracking, super-critical treat-ment, and steam reforming [14]. Catalytic cracking can onlypartially deoxygenate the raw bio-oil and produces low liquidyields [14]. Steam reforming produces a low energy densitygaseous fuel and supercritical treatment requires high capitalcost due to the requirement of high pressure vessels [15].Other upgrading methods such as olefination and esterification[16, 17] are used to upgrade bio-oil to boiler fuel.

HDO is one of the upgrading methods reported to pro-duce a highly de-oxygenated (as low as zero wt % oxygen inthe upgraded product), high energy liquid fuel. PreviousHDO studies [4, 18, 19] over nearly three decades have pro-vided considerable information about methods to upgradebio-oil by this technology. However, rapid catalyst deactiva-tion (by coking), reactor plugging, and low product yieldscontinue to be problematic and further research is requiredto refine current HDO methods and catalysts. Our studiesemployed a pretreated bio-oil (PTBO) treated by HDOmethod to upgrade bio-oil to mixed liquid hydrocarbons.Moreover, instead of employing a two-step upgradingmethod (hydrotreatment followed by hydrocracking), asingle-stage hydroprocessing was applied to produce liquidfuel from bio-oil.

It has become customary to practice the hydroprocessingof bio-oil by utilizing a 2-stage approach in which the 1ststage comprises a hydrotreating stage utilizing a mild temper-ature (<300�C) for the reaction. This 1st stage eliminates thepolymerization of bio-oil that occurs when raw bio-oil issubjected to high temperatures. Hydrocracking the hydro-treated product is then practiced in a 2nd stage reaction athigher temperature (>350�C). The 2-stage hydroprocessingmethod usually requires two reactors which increases

Presented at the Tcbiomass2013 conference; September 3–6, 2013,at Chicago, Illinois.

VC 2014 American Institute of Chemical Engineers

Environmental Progress & Sustainable Energy (Vol.33, No.3) DOI 10.1002/ep726 October 2014

Page 2: Single stage hydroprocessing of pyrolysis oil in a continuous packed-bed reactor

the capital cost of the hydroprocessing technology; morereaction time is also required increasing variable costs.The objective of this study was to apply a single-stage hydro-processing treatment to upgrade a PTBO to a hydrocarbonmixture.

MATERIALS AND METHODS

FeedstockPTBO was used as a feedstock for the hydrotreatment

process. A patent application [20] has been filed to protectthe intellectual property represented by the production ofPTBO. The patent application has been recently published[20]. Raw bio-oil was produced using an augur pyrolysisreactor operated at a temperature between 400 and 550�C.PTBO was produced by mixing the raw bio-oil. Followingthe addition of oxone and hydrogen peroxide the mixturewas stirred for 90 min at room temperature and ambientpressure. Following this treatment the patent embodimentwhereby butyric anhydride was added [20] was performedfor this study. For this step, the mixture was heated at 90�Cat ambient pressure to obtain the final PTBO. The pretreat-ment step, unlike other bio-oil mild hydrotreating processes,was performed at low temperatures (below 100�C) and atambient pressure (no hydrogen pressure) in a Parr autoclave(450 mL) (Parr instruments and Co).

Catalysts

CoMo/g-Al2O3 and Ru/g-Al2O3 were purchased commer-cially from Alfa Aesar and Acros Organics, respectively. Cata-lyst supports such as Si-Al, g-alumina, and the requiredinorganic metal salts for preparation of Ni/Si-Al and FeW/Si-Al catalysts were also commercially purchased. The Ni/Si-Aland FeW/Si-Al catalysts were prepared using the wet-impregnation method, whereby the metal salts were impreg-nated on catalyst supports and then dried at 120�C for 4–6 hbefore being calcined at 550�C for 4 h. The calcined metal-dispersed catalysts were then reduced at 700�C using hydro-gen flow (100 mLpm) for 4 h. For the 1:2 ratio of Ru/g-Al2O3

to Ni/Si-Al catalyst the Ru/g-Al2O3 and Ni/Si-Al were loadedin separate heating zones of the continuous packed bedreactor. In the current study only reduced catalysts weretested. For all catalysts physical properties such as acid value(AV), water content (H2O%), higher heating value (HHV),and oxygen content were compared. Based on the catalystthat performed the best in terms of physical properties, fur-ther tests on that catalyst were performed. These tests weresimulated distillation (SIMDIS) and detailed hydrocarbonanalysis (DHA). The pressure drop was measured in 2 hintervals over an 8 h period on time on stream for each cata-lyst in the packed-bed reactor as shown in Table 1.

Continuous Packed-Bed ReactorThe continuous packed-bed reactor (Figure 1) consisted

of a 100 I.D tubular reactor enclosed in a three-zone furnace(three 600 zones independently controlled by three tempera-ture controllers) followed by a condensation system. Thetemperatures inside the reactor were monitored with a pointprofile thermocouple with ten sensing points (Omega Instru-ments). The catalyst bed temperature was maintained at adesired set point (shown in Table 1) through the course ofthe experiments. The bio-oil was fed using a high pressuredual-pump system (Teledyne Isco 500D) and the hydrogenflow was controlled by a mass flow controller (MFC; BrooksInstruments). The reactor pressure was controlled using aback-pressure regulator.

The reactor was loaded with catalyst and initially set to150�C. Once initial temperature set point was attained, thereactor temperature was raised in 100�C intervals to the

desired reaction temperature of 350–375�C. (Over the courseof the reaction the temperature from Zones 1–3 fluctuatedbetween 300 and 375�C from adiabatic conditions producedby the catalytic reaction.) The reactor was then pressurizedto the desired hydrogen reaction pressure using a by-passvalve. Once the system reached the desired pressure of 1500psi, hydrogen was supplied to the reactor through a massflow controller set at the desired flow rate (0.5 L/min).

All experiments were carried out at an LHSV of 0.2–0.3h21. The exit gas flow rate was monitored using an Agilentflow meter. Products exiting from the packed-bed reactorwere cooled in the condenser and the liquid products werecollected in a sampling bottle at regular intervals, while gassampling was performed using Tedlar sampling bags.

Physical and Chemical AnalysisPTBO and the liquid products produced from the HDO

treatments were analyzed for AV, H2O%, HHV, elementalanalysis, and gas chromatography - mass spectrometry (GC-MS) analysis. AV was obtained by dissolving one gram ofbio-oil in an isopropanol/water mixture and titrating to a pHof 8.5 with 0.1 N KOH by ASTM D 664. Percent water wasdetermined by ASTM Method E 203 with a Cole-ParmerModel C-25800-10 titration apparatus. The HHV was deter-mined with a Parr 6400 automatic isoperibol calorimeter byASTM D 240. Elemental analyses of raw bio-oil were per-formed with an EA920 elemental analyzer by ASTM D 5291.DHA was performed using ASTM D6730-01 method. DHAwas performed in the PerkinElmer Clarus 680 GC equippedwith a built-in model Arnel 4060 Detailed Hydrocarbon Ana-lyzer. SIMDIS was performed by the ASTM D2887 method ona gas chromatograph.

RESULTS AND DISCUSSION

Our study focused on testing the efficacy of various cata-lysts in single-stage hydroprocessing of PTBO to liquidhydrocarbons. The catalysts employed for this study includedFeW/Si-Al, CoMo/g-Al2O3, iron oxide-chromium oxide (Fe-Cr) catalyst, and 1:2 ratio of Ru/g-Al2O3:Ni/Si-Al. The col-lected liquid products were centrifuged for 1 h to separatethe aqueous and hydrocarbon fractions. The hydrocarbonfraction of the liquid products were subjected to character-izations such as elemental analysis, HHV, AV, and H2O%analysis to determine quality, based on which the catalystperformance was interpreted. The reduction in AV, increase

Table 1. Effect of pressure drop of packed-bed reactor bycatalyst applied with time on stream.

Catalyst

Pressuredifference

(psi)Time(h)

FeW/Si-Al 3 23 44 64 8

CoMo/ g-Al2O3 3 23 44 64 8

Fe-Cr mixed oxide 3 22 44 65 8

1:2 ratio of Ru/g-Al2O3

and Ni/Si-Al2 23 44 63 8

Environmental Progress & Sustainable Energy (Vol.33, No.3) DOI 10.1002/ep October 2014 727

Page 3: Single stage hydroprocessing of pyrolysis oil in a continuous packed-bed reactor

in HHV accompanied by decreased elemental oxygen in thehydrocarbon fraction of the product indicated reduction ofbio-oil oxygen. The pressure drop values were measuredwith time on stream for 8 h in a packed-bed reactor asshown in Table 1.

Table 2 compares PTBO properties to those of four testedcatalysts: FeW/Si-Al, CoMo/g-Al2O3, Fe-Cr mixed oxide cata-lyst, and 1:2 ratio of Ru/g-Al2O3, Ni/Si-Al. Applying these cat-alysts to the PTBO hydroprocessing produced respective AVsof 51.9, 2.1, 24.8, and 16.5 mg of KOH/g oil. As evident from

Table 2, all the catalysts significantly decreased the AV, withCoMo/ g-Al2O3 being the most effective catalyst. HHVs pro-duced by these catalysts were 33.0, 43.1, 35.4, and 41.6 MJ/kg, respectively. Table 3 shows the elemental oxygen in theliquid products obtained using the above mentioned catalystswere 7.9%, 0.1%, 19.5%, and 0.6%, respectively. The bestresults in terms of AV, HHV, H2O%, and elemental oxygenwere obtained using the CoMo/g-Al2O3 catalyst and were2.14, 43.1, 0.2, and 0.1%, respectively.

All liquid products obtained from the HDO experimentscontained two (aqueous and hydrocarbon) fractions. Figure2, showing the H2O% content of the hydrocarbon fractions,indicated that the CoMo/g-Al2O3 resulted in the formation ofa hydrocarbon fraction with negligible (0.2%) water content.The 1:2 ratio of Ru/g-Al2O3 and Ni/Si-Al, and Fe-Cr, FeWupgraded hydrocarbon fractions showed respective watercontents of 0.3, 4.9, and 7.6%. Figure 2 shows that AV fol-lowed the trend of H2O% produced by each catalyst. Thisresult indicates that higher H2O% values dilute AV and viceversa.

Figure 3 shows the effect of hydrotreating catalysts inincreasing the HHV of the hydrocarbon fractions. All hydro-carbon fractions showed increased HHVs compared to thePTBO. While CoMo/g-Al2O3 was the most effective catalystin improving the HHV of the hydrocarbon fraction (�43 MJ/

Figure 1. Schematic of the continuous packed-bed reactor.

Table 2. Hydrotreating catalysts, experimental conditions applied, and properties (AV, HHV, H2O%) of the hydrocarbon frac-tion of the liquid products.

CatalystTemperature

(�C) P (Psig) LHSV (h21) AV HHV H2O%

Control-Pretreated bio-oil (PTBO) 223.0 19.0 21.0FeW/Si-Al 300–375 1500 0.2–0.3 51.9 33.0 7.6CoMo/g-Al2O3 2.1 43.1 0.2Fe-Cr mixed oxide 24.8 35.4 4.91:2 ratio of Ru/g-Al2O3 and Ni/Si-Al 16.5 41.6 0.3

Table 3. The elemental analysis of the hydrocarbon fractionproduced by catalyst type.

Catalyst %C %H2 %N2

%O2

(100 2 %C1 %H2 1 %N2)

PTBO 43.1 8.9 0.2 47.8FeW/Si-Al 81.7 10.0 0.4 7.9CoMo/g-Al2O3 86.3 13.4 0.4 0.1Fe-Cr mixed oxide 69.1 11.2 0.2 19.51:2 ratio of Ru/g-Al2O3

and B: Ni/ Si-Al85.5 13.8 0.1 0.6

Environmental Progress & Sustainable Energy (Vol.33, No.3) DOI 10.1002/ep728 October 2014

Page 4: Single stage hydroprocessing of pyrolysis oil in a continuous packed-bed reactor

kg), 1:2 ratio of Ru/g-Al2O3, Ni/Si-Al was also effective inincreasing the HHV (�42). From Figures 3 and 4 it can bedetermined that the increase in HHV of the hydrocarbonfractions was a function of the decrease in the water contentand AVs.

As shown in Table 4, the total yields produced by FeW/Si-Al, CoMo/g-Al2O3, Fe-Cr mixed oxide, and 1:2 ratio of Ru/g-Al2O3 and Ni/Si-Al catalyst were 0.34, 0.34, 0.44, 0.47 g/g

of feed, respectively, for a time on stream of 8 h. Among thetested catalysts, the highest total liquid yields were producedby Fe-Cr mixed oxide catalyst and 1:2 ratio of Ru/g-Al2O3,Ni/Si-Al catalysts, but the overall quality of the product prop-erties deteriorated over 8 h stream with these catalysts.Though the total liquid yields were less with CoMo/g-Al2O3

catalyst, the properties of the product obtained did not differover the 8 h on stream. Table 5 gives the DHA of mixed liq-uid hydrocarbons obtained by the catalysts indicating in fig-ures 3 and 4 to produce the highest HHVs: CoMo/g-Al2O3

and 1:2 ratio of Ru/g-Al2O3, Ni/Si-Al catalyst. For the reducedCoMo/g-Al2O3 catalyst, the olefins, iso-paraffins, and C141compounds dominated the mixed liquid hydrocarbons fol-lowed by naphthenes and aromatics with an octane value of52.33 on an average of three replicates. In comparison, theDHA of the 1:2 ratio of Ru/g-Al2O3, Ni/Si-Al catalystupgraded product predominantly contained iso-paraffins,olefins, and aromatics over C141 compounds, naphthenesand iso-paraffins with an octane value of 50.20 on an aver-age of three replicates.

Table 6 shows the analysis of gas samples collected dur-ing the hydroprocessing experiments. Gas samples were col-lected 10–15 min prior to collecting the liquid samples. Gasanalysis was used to interpret the H2 percentage in the exitgas as well as the formation of low molecular weight hydro-carbon gases such as CH4, C2H6, etc. As seen below in Table6, exit gases from the CoMo/g-Al2O3 experiment containedthe lowest hydrogen percentage, indicating that more hydro-gen was consumed for hydroprocessing PTBO. This observa-tion is in agreement with the elemental analysis results thatshowed the presence of 13.4% hydrogen in the CoMo/g-Al2O3 upgraded oil fraction compared to 10.6% and 11.2 %hydrogen present in the FeW/Si-Al and Fe-Cr mixed oxideupgraded hydrocarbon fractions. The exit gases from theexperiment using Ru/g-Al2O3 1 Ni/Si-Al catalyst combinationhad 74% hydrogen which had the next best performance tothe CoMo/g-Al2O3 gas samples.

Figure 2. Effect of catalysts on H2O% of the liquid products.[Color figure can be viewed in the online issue, which isavailable at wileyonlinelibrary.com.]

Figure 3. Effect of catalysts on AV and HHV of the liquidproducts. [Color figure can be viewed in the online issue,which is available at wileyonlinelibrary.com.]

Figure 4. Elemental oxygen in the liquid products obtainedpost hydroprocessing. [Color figure can be viewed in theonline issue, which is available at wileyonlinelibrary.com.]

Table 4. Effect of catalyst type on total liquid, hydrocarbonfraction (HF) and aqueous fraction (AF) yields (based on dryweight of biomass).

Catalyst

Yields (g/g of feed)

Totalliquid HF AF

FeW/Si-Al 0.34 0.05 0.29CoMo/g-Al2O3 0.34 0.02 0.32Fe-Cr mixed oxide 0.44 0.10 0.341:2 ratio of Ru/g-Al2O3

and Ni/Si-Al0.47 0.14 0.33

Table 5. DHA mass percentage of CoMo/g-Al2O3 and 1:2ratio of Ru/g-Al2O3 and Ni/Si-Al upgraded oil fractions(ASTM D6730-01).

Catalyst CoMo/c-Al2O3

1:2 ratio of Ru/c-Al2O3 and Ni/Si-Al

Paraffins 2.98 0.45I-Paraffins 18.19 23.90Olefins 17.51 14.83Naphthenes 14.18 9.45Aromatics 7.97 11.22Total C141 13 10.90Unknowns 25.58 28.19Octane no. 52.33 50.20

Environmental Progress & Sustainable Energy (Vol.33, No.3) DOI 10.1002/ep October 2014 729

Page 5: Single stage hydroprocessing of pyrolysis oil in a continuous packed-bed reactor

SIMDIS was perfomed according to ASTM D2887 methodon the CoMo/g-Al2O3 upgraded liquid fuel and the plot isshown in Figure 5. SIMDIS of the fuel showed initial boilingpoint (IBP) and final boiling point (FBP) to be 69�C and304�C, respectively. Therefore, it can be reported that theobtained liquid fuel may have a boiling temperature range of69–304�C and contain petroleum equivalents of 50% gasoline(38–170�C), 30% jet fuel (170–250�C), and 20% diesel (250–304�C) range hydrocarbons. SIMDIS showed no presenceof vacuum gas oil (VGO) range (> 315�C) hydrocarbons inour fuel.

SUMMARY

This research successfully demonstrated that PTBO can behydrotreated to a 100% hydrocarbon mixture utilizing onlysingle-stage HDO (hydrotreating and hydrocracking) in apacked-bed reactor. Among the catalysts tested for thissingle-stage experiment, the best results were obtained usingreduced CoMo/g-Al2O3 catalyst for HDO of the PTBO at300–375�C with a hydrogen flow rate of 0.5 L/min at 1500Psig H2 pressure and an LHSV of 0.2–0.3 h21. The obtainedliquid hydrocarbon had an AV of 2.14 with a heating value of43.1 MJ/kg, elemental oxygen of 0.06%, and water content of0.18%. Gas analysis results indicated that the reduced CoMo/g-Al2O3 catalyst had the highest hydrogen consumptionamong all the catalysts. SIMDIS results showed that the liquidhydrocarbon contained petroleum equivalents of 50% gaso-line (38–170�C), 30% jet fuel (170–250�C), and 20% diesel(250–304�C) range hydrocarbons. The high activity of pro-moted Mo catalyst (CoMo/g-Al2O3) can be attributed to theavailability of number of “d” electrons in the highest occu-pied orbitals. The promoter element (Co) aids in reducingthe oxidation state of the Mo atom by donating electrons toMo atoms and thereby increasing the number of 4delectrons.

FUTURE STUDY

Our current study indicated that PTBO can be upgradedvia HDO in a single-stage treatment. However, improvementin liquid yields is highly desirable. Catalyst screening studies,including sulfided CoMo, will be continued to attempt toimprove liquid hydrocarbon yields. In addition, longer-termdeactivation studies will be performed to examine the effi-cacy of the best performing catalysts in extended runs.

ACKNOWLEDGMENTS

Authors gratefully acknowledge the financial support pro-vided by the Sustainable Energy Research Center at Missis-sippi State University that is supported by the Department ofEnergy under Award Number DE-FG3606GO86025.

DISCLAIMER

This report was prepared as an account of work spon-sored by an agency of the United States government. Neitherthe United States Government nor any agency thereof, norany of their employees, makes any warranty, express orimplied, or assumes any legal liability or responsibility forthe accuracy, completeness, or usefulness of any informa-tion, apparatus, product, or process disclosed, or representsthat its use would not infringe privately owned rights. Refer-ence herein to any specific commercial product, process, orservice by trade name, trademark, manufacturer, or otherwisedoes not necessarily constitute or imply its endorsement, rec-ommendation, or favoring by the United States Governmentor any agency thereof. The views and opinions of authorsexpressed herein do not necessarily state or reflect those ofthe United States Government or any agency thereof.

LITERATURE CITED

1. Carlson, T. R., Tompsett, G. A., Conner, W. C., & Huber,G. W. (2009). Aromatic production from catalytic fastpyrolysis of biomass-derived feedstocks, Topics in Cataly-sis, 52, 241–252.

2. Demiral, L., & Sensoz, S. (2008). Pyrolysis of industrialwastes (olive and hazelnut bagasse), Bioresource Tech-nology, 99, 8002–8007.

3. Mohan, D., Pittman, C.U., Jr., & Steele, P.H. (2006). Pyrol-ysis of wood/biomass for bio-oil: A critical review,Energy & Fuels, 20, 848–889.

4. Elliott, D.C. (2007). Historical development in hydropro-cessing bio-oils, Energy & Fuels, 21, 1792–1815.

5. Furimsky, E. (2000). Catalytic hydrodeoxygenation,Applied Catalysis A: General, 199, 147–190.

6. Marcelo, E., Domine, C., Veen, V., Schuurman, Y., &Mirodatos, C. (2008). Coprocessing of oxygenated bio-mass compounds and hydrocarbons for the productionof sustainable fuel, ChemSusChem, 1, 179–181.

7. Zhang, Q., Chang, J., Wang, T., & Xu, Y. (2006). Upgrad-ing bio-oil over different solid catalysts, Energy Fuels, 20,2717–2720.

8. Diebold, J.P. (2000). Review of the chemical and physicalmechanisms of the storage stability of fast pyrolysis bio-oils, Golden, CO: National Renewable Energy Laboratory.2000; Report No. NERL/SR-570–27613. Also available at:http://www.osti.gov/bridge.

Table 6. Gas analysis results from hydroprocessing experiments by catalyst type.

Catalyst H2% O2% N2% CH4% CO% CO2% C2H6%

FeW/Si-Al 79.21 0.41 1.36 0.06 0.44 1.01 0.02CoMo/g-Al2O3 69.79 0.25 0.72 5.73 0.16 1.66 0.93Fe-Cr 81.74 0.48 1.5 0.25 0 1.84 0.06Ru/g-Al2O3 1 Ni/Si-Al 74.38 0.15 0.38 7.23 0 0.36 0.29

Figure 5. Weight % distilled vs boiling temperature Tb (�C) ofthe CoMo/g-Al2O3 upgraded liquid fuel determined by SIMDISGC (ASTM D2887). [Color figure can be viewed in the onlineissue, which is available at wileyonlinelibrary.com.]

Environmental Progress & Sustainable Energy (Vol.33, No.3) DOI 10.1002/ep730 October 2014

Page 6: Single stage hydroprocessing of pyrolysis oil in a continuous packed-bed reactor

9. Wildschut, J., Mahfud, F.H., Venderbosch, R.H., & Heeres,H.J. (2009). Hydrotreatment of fast pyrolysis oil usingheterogeneous noble metal catalysts, Industrial & Engi-neering Chemistry Research, 48, 10324–10334.

10. Senol, O.I., Viljava, T.R., & Krause, A.O. (2005). Hydro-deoxygenation of methyl esters on sulphided NiMo/gamma-Al2O3 and CoMo/gamma-Al2O3 catalysts, Cataly-sis Today, 100, 331–335.

11. De Miguel Mercader, F., Groeneveld, M.J., Kersten, S.R.A.,Way, N.W.J., Schaverien, C.J., & Hogendoorn, J.A. (2010).Production of advanced biofuels: Co-processing ofupgraded pyrolysis oil in standard refinery units, AppliedCatalysis B: Environmental, 96, 57–66.

12. Gald�amez, J.R., Garc�ıa, L., & Bilbao, R. (2005). Hydrogenproduction by steam reforming of bio-oil using coprecipi-tated Ni-Al catalysts. Acetic acid as a model compound,Energy & Fuels, 19, 1133–1142.

13. Wang, D., Montane, D., & Chornet, E. (1996). Catalyticsteam reforming of biomass-derived oxygenates: Aceticacid and hydroxyacetaldehyde, Applied Catalysis A: Gen-eral, 143, 245–270.

14. Wang, Z.X., Pan, Y., Dong, T., Zhu, X.F., Kan, T., Yuan,L.X., Torimoto, Y., Sadakata, M., & Li, Q.X. (2007). Pro-duction of hydrogen from catalytic steam reforming ofbio-oil using C12A7-O-based catalysts, Applied CatalysisA General, 320, 24–34.

15. Vadillo,V., Sanchez-Oneto, J., Portela, J.R., & Martinez dela Ossa, E. (2013). Problems in supercritical water oxida-tion process and proposed solutions, Industrial & Engi-neering Chemical Research, 52, 7617–7629.

16. Mahfud, F.H., Melian-Cabrera, I., Manurung, R., &Heeres, H.J. (2007). Biomass to fuels—Upgrading of flashpyrolysis oil by reactive distillation using a high boilingalcohol and acid catalysts, Process Safety and Environ-mental Protection, 85, 466–472.

17. Zhang, Q., Chang, J., Wang, T.J., & Xu, Y. (2006). Prepara-tion of solid acid SO42-/SiO2-TiO2 and its catalytic activityfor esterification, Chinese Journal of Catalysis, 27, 1033–1038.

18. Elliott, D.C., Hart, T.R., Neuenschwander, G.G., Rotness,L.J., & Zacher, A.H. (2009). Catalytic hydroprocessing ofbiomass fast pyrolysis bio-oil to produce hydrocarbonproducts, Environmental Progress & Sustainable Energy,28, 441–449.

19. Elliott, D.C., Hart, T.R., Neuenschwander, G.G., Rotness,L.J., Mariefel, V., Zacher, A.H., and Solantauta, Y. (2012).Catalytic hydroprocessing of fast pyrolysis bio-oil frompine sawdust, Energy & Fuels, 26, 3891–3896.

20. Steele, P., Gajjela, S., & Tanneru, S. (2013). Compositionand methods for improved fuel production, US patentpublication no. US 20130291431 A1, Nov 7.

Environmental Progress & Sustainable Energy (Vol.33, No.3) DOI 10.1002/ep October 2014 731