Historical Developments in Hydroprocessing Bio-Oils

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 24

1792 Energy & Fuels 2007, 21, 1792-1815

Historical Developments in Hydroprocessing Bio-oils


Douglas C. Elliott*
Pacific Northwest National Laboratory, P.O. Box 999, 902 Battelle BouleVard,
Richland, Washington 99352

ReceiVed January 25, 2007. ReVised Manuscript ReceiVed March 16, 2007

This paper is a review of the developments in the field of catalytic hydroprocessing of biomass-derived
liquefaction conversion products (bio-oil) over the past 25 years. Work has been underway, primarily in the
U.S. and Europe, in catalytic hydrotreating and hydrocracking of bio-oil in both batch-fed and continuous-
flow bench-scale reactor systems. A range of heterogeneous catalyst materials have been tested, including
conventional sulfided catalysts developed for petroleum hydroprocessing and precious metal catalysts. The
important processing differences have been identified, which required adjustments to conventional hydropro-
cessing as applied to petroleum feedstocks. This application of hydroprocessing is seen as an extension of
petroleum processing and system requirements are not far outside the range of conventional hydroprocessing.
The technology is still under development but can play a significant role in supplementing increasingly expensive
petroleum.

Introduction data from the experimentation of handling and upgrading bio-


oils by catalytic hydrotreating.
The objective of this review was to gather the relevant
information from the literature (which is in specific cases more Biomass Liquefaction Processes
or less accessible from earlier times) to describe the development
of hydrotreating upgrading technology to produce transportation By way of introduction, the thermochemical conversion
liquid fuels from thermochemically derived biomass liquids. methods for producing oil from biomass are reviewed. There
This paper provides only an overview of the types of technology are basically two methodsshigh-pressure liquefaction (typically
for liquid fuels production from biomass, such as hydrothermal hydrothermal for biomass) and atmospheric-pressure fast py-
processing or fast pyrolysis. The important consideration is that rolysis.
these liquid products from biomass are not useful as fuels other Hydrothermal Processes. As early as the 1920s, experi-
than direct boiler firing and possibly for some types of turbine mental data from Berl supported the concept of oil production
and large diesel applications after significant modifications. In from biomass in hot water using alkali as catalyst.4 Heinemann
order for the biomass liquids to be useful as transportation fuels, continued this work into the 1950s.5 Following the Arab oil
they require chemical transformation to increase volatility and embargo of 1974, initial efforts in biomass liquefaction in the
thermal stability and reduce viscosity through oxygen removal U.S. focused on the high-pressure (hydrothermal) liquefaction
and molecular weight reduction. The reader may refer to process based on this concept developed at the Pittsburgh Energy
Bridgwater et al. for a discussion of how these types of Research Center (PERC) and demonstrated at the Albany
technologies relate to each other.1 Maggi and Delmon provide Biomass Liquefaction Experimental Facility at Albany, Oregon,6
a detailed comparison of fast and slow pyrolysis.2 The field of at a scale of 100 kg/h. In this process, wood powder was slurried
hydrotreating of hydrothermal (high-pressure) liquefaction oils in recycle oil product and mixed with water and a sodium
from biomass provides the initial basis for this review. This carbonate catalyst. The slurry was pumped to a pressurized
study then focuses on the hydroprocessing of bio-oil from fast reactor where carbon monoxide was added at 20.8 MPa. The
pyrolysis to liquid fuels. As such, it does not provide details of slurry was held at a temperature of around 350 °C for 20 min
the liquefaction processes themselves but will focus on the to 1 h after which it was depressurized, and oil product was
catalytic processes used to convert the highly oxygenated bio- separated from water. This heavy oil product had a melting point
oil to a liquid more similar to petroleum-derived fuels. near room temperature with an oxygen content of 12 to 14%
This review included collection of data from the literature in and dissolved water of 3 to 5%.
the field of study highlighting developments from the 1980s to A later derivative of this process was the HTU process from
the present. A useful summary of the early work has been Shell.7 In this process, similar conditions were used but the
published.3 This review is a collection and summary of the actual
(4) Berl, E. Science 1944, 99, 309.
(5) Heinemann, H. Hydrocarbons from Cellulosic Wastes. Pet. Refiner
* E-mail address: dougc.elliott@pnl.gov. 1954, 33 (7), 161-163.
(1) Bridgwater, A. V.; Meier, D.; Radlein, D. An Overview of Fast (6) Thigpen, P. L.; Berry, W. L. In Energy from Biomass and Wastes
Pyrolysis of Biomass. Org. Geochem. 1999, 30, 1479-1493. VI; Klass, D.L., Ed.; Institute of Gas Technology: Chicago, 1982; p 1057.
(2) Maggi, R.; Delmon, B. Comparison between “Slow” and “Flash” (7) Goudriaan, F.; van de Beld, B.; Boerefijn, F. R.; Bos, G. M.; Naber,
Pyrolysis Oils from Biomass. Fuel 1994, 73 (5), 671-677. J. E.; van der Wal, S.; Zeevalkink, J. A. Thermal Efficiency of the HTU
(3) Elliott, D. C.; Beckman, D.; Bridgwater, A. V.; Diebold, J. P.; Gevert, Process for Biomass Liquefaction. In Progress in Thermochemical Biomass
S. B.; Solantausta, Y. Developments in Direct Thermochemical Liquefaction ConVersion; Bridgwater, A.V., Ed.; Blackwell Science, LTD.: Oxford,
of Biomass: 1983-1990. Energy Fuels 1991, 5 (3), 399-410. England, 2001; pp 1312-1325.

10.1021/ef070044u CCC: $37.00 © 2007 American Chemical Society


Published on Web 05/02/2007
DeVelopments in Hydroprocessing Bio-oils Energy & Fuels, Vol. 21, No. 3, 2007 1793

reducing agent alkali/CO was left out; the resulting product had have been wide-ranging efforts reported in the literature, as a
an oxygen content nearer to 18% with a melting point of about recent review describes.12 A large portion of the body of work
80 °C.8 This process has been scaled up to a 10 kg/h dry biomass addresses the catalytic chemistry by hydrotreating model
feed pilot plant. The pilot plant was operated for a 500 h design compounds containing oxygen. Many of these models are
run in 2004. relevant to bio-oil hydroprocessing, such as phenolics and
Fast Pyrolysis Processes. Most of the work up to the year aromatic ethers. This report provides a detailed review of those
2000 is covered in detail in a review article by Bridgwater and research efforts that focused on processing actual bio-oil
Peacocke,9 but some of the highlights are described below. products, and model compound studies are addressed only when
Modern developments in fast pyrolysis can be traced to the performed as a part of the bio-oil upgrading research effort.
Occidental/Garrett process development work in the 1970s. That
effort focused on waste processing, but a detailed report exists Hydrothermal Liquefaction Product Upgrading
which describes results with biomass materials.10 The University
As the first wood liquefaction oil produced at pilot scale
of Waterloo first published results in a fluidized bed for fast
became available in 1979, oil upgrading process research began.
pyrolysis of biomass in a government report in 1981. After
The catalytic hydrotreating of the heavy oil product followed
extensive research, the technology was scaled up in Spain
the methods used in conventional petroleum processing technol-
(Union Fenosa) and is now a patented technology held by
ogy.
Dynamotive Energy Systems Corporation. A company which
Pacific Northwest National Laboratory (PNL/PNNL). Bio-
grew out of the University of Waterloo, Resource Transforms
oil upgrading work at PNNL has focused on heterogeneous
International Ltd., is still developing refinements and chemical
catalytic hydroprocessing. Initial work involved batch reactor
products from fast pyrolysis.
tests of model phenolic compounds13 with various catalysts.
A second version of fast pyrolysis which utilizes circulating
Commercial samples of catalysts were used representing CoMo,
fluidized beds was developed out of the University of Western
NiMo, NiW, Ni, Co, Pd, and CuCrO to hydrogenate phenol at
Ontario and is now commercialized by Ensyn Technologies
300 or 400 °C (1 h at temperature). P-Cresol, ethyl-phenol,
(RTP, rapid thermal processing). Their technology is used in
dimethyl-phenol, trimethyl-phenol, naphthol, and guaiacol (meth-
food flavoring production plants in the U.S. It was scaled up
oxy-phenol) were also tested with a CoMo catalyst at 400 °C.
for further development in Italy in the 1990s but was not
Of the catalysts tested, the sulfided form of CoMo was most
operated extensively.
active, producing a product containing 33.8% benzene and 3.6%
Other versions of fast pyrolysis include the ablative reactor cyclohexane at 400 °C. The Ni catalyst was also active,
first demonstrated in its vortex form by the National Renewable producing a product with 16.9% benzene and 7.6% cyclohexane.
Energy Laboratory (formerly the Solar Energy Research Insti- The sulfided Ni catalyst still produced 8.0% cyclohexane, but
tute) and now being developed at Aston University and Twente its yield of benzene dropped to near zero (0.4%). A Pd catalyst
University with rotating plate and cone reactors. The Twente produced a 7.8% benzene product with 2.7% cyclohexane but
version has been scaled up in Malaysia. The simple entrained 5.5% cyclohexanone. At lower temperature (300 °C), cyclo-
flow reactor was attempted at the Georgia Tech Research hexanone was the primary product at 8.1% and benzene and
Institute in the 1980s as an extrapolation from the Tech-Air cyclohexane were nearly equal at 2.0 and 2.5%, respectively.
upflow pyrolysis/gasification technology. The operations could The original work with hydrotreating biomass-derived liquids
never achieve as high oil yields as those found in the other was the effort to make gasoline from the high-pressure liquefac-
reactors, apparently because of the limited heat transfer. Vacuum tion oil produced from wood at the Albany Biomass Liquefac-
pyrolysis was developed to demonstration scale at Laval tion Pilot Plant. The oil from high-pressure liquefaction (a
University but was not found to be economical with biomass. hydrothermal, alkali-catalyzed process) was a more deoxygen-
ated product than fast pyrolysis bio-oil. As a result, it was more
Bio-oil Hydroprocessing thermally stable and required less hydrogenation to produce a
gasoline product. Processing technology was adapted from
Upgrading biomass-derived oils to hydrocarbon fuels requires
conventional petroleum hydrotreating using nickel and sulfided
oxygen removal and molecular weight reduction. As a result,
cobalt-molybdenum catalysts14 in a continuous-flow, fixed
there is typically a formation of an oil phase product and a
catalyst bed reactor system operated in an upflow configuration.
separate aqueous phase product by hydroprocessing. To mini-
The catalysts used in these tests are identified and described
mize hydrogen consumption in hydroprocessing, hydrodeoxy-
in Table 1.
genation (HDO) must be emphasized, without saturation of the
Preliminary results with a light oil fraction showed that the
aromatic rings. Hydroprocessing of biomass-derived oils differs
sulfided form of the CoMo catalyst was much more active than
from processing petroleum or coal liquids because of the
the oxide form. A copper-chromite catalyst was also tested
importance of deoxygenation, as opposed to nitrogen or sulfur
and found to be even less active. The results are presented in
removal. At the time that research began to evaluate the HDO
Table 2 for extinction hydrotreating/hydrocracking of whole
of biomass-derived oils, HDO had received only limited
wood oil and a distillate product. The whole oil could be
attention in the literature.11 Over the past 20 plus years, there
hydrotreated much like the distillate oil but required a lower
space velocity and operating pressure and resulted in higher
(8) Goudriaan, F.; Zeevalkink, J. A.; Naber, J. E. HTU Process Design
and Development: Innovation Involves Many Disciplines. In Science in hydrogen consumption. The nickel catalyst exhibited activity
Thermal and Chemical Biomass ConVersion; Bridgwater, A. V., Boocock,
D. G. B., Eds.; CPL Press: Newbury Berks, U.K., 2006; pp 1069-1081. (12) Furimsky, E. Catalytic Hydrodeoxygenation. Appl. Catal., A: Gen.
(9) Bridgwater, A. V.; Peacocke, G. V. C. Fast Pyrolysis Processes for 2000, 199, 147-190.
Biomass. Renewable Sustainable Energy ReV. 2000, 4, 1-73. (13) Elliott, D. C. Hydrodeoxygenation of Phenolic Components of
(10) Boucher, F. B.; Knell, E. W.; Preston, G. T.; Mallan, G. M. Pyrolysis Wood-Derived Oil. Prepr. Pap.sAm. Chem. Soc., DiV. Pet. Chem. 1983,
of Industrial Wastes for Oil and ActiVated Carbon RecoVery; report no. 28 (3), 667-674.
EPA-600/2-77-091, project no. S-801202, U.S. EPA: Washington, D.C., (14) Elliott, D. C.; Baker, E. G. Upgrading Biomass Liquefaction
May 1977. Products through Hydrodeoxygenation. Biotechnol. Bioeng. Symp. 1984,
(11) Furimsky, E. Catal. ReV.sSci. Eng. 1983, 25 (3), 42. 14, 159-174.
1794 Energy & Fuels, Vol. 21, No. 3, 2007 Elliott

Table 1. Catalysts Used in Hydrotreating/Hydrocracking Tests


supplier catalyst id active metals weight % support forma
Harshaw Ni-1404 Ni 68 proprietary 1/8-in T
Harshaw CoMo0402 CoO/MoO3 3/15 silica-alumina 1/8-in T
Harshaw HT 400 CoO/MoO3 3/15 γ-Al2O3 1/8-in E
Harshaw HT 500 NiO/MoO3 3.5/15.5 γ-Al2O3 1/8-in E
Harshaw Ni-3266 Ni 50 silica-alumina 1/16-in E
Harshaw Ni-4301 NiO/WO2 6/19 γ-Al2O3
Haldor Topsoeb TK 710 CoO/MoO3 2/6 Al2O3 3/16-in R
Haldor Topsoeb TK 750 CoO/MoO3 2.3/10 Al2O3 1/16-in E
Haldor Topsoeb TK 770 CoO/MoO3 3.4/14 Al2O3 1/16-in E
Haldor Topsoe TK 751 NiO/MoO3 3/13 Al2O3 1 mm E
Katalco CoMo479 CoO/MoO3 4.4/19 Al2O3 1/16-in E
Katalco CoMo499 CoO/MoO3 4.4/19 γ-Al2O3 1/16-in E
Katalco KAT 4000 CoO/MoO3 3.5/14 γ-Al2O3 1/32-in E
Shell S411 NiO/MoO3 2.67/14.48 Al2O3 c 1/20-in T
PNL/Linde CoMo/Y CoO/MoO3 3.5/13.9 Y-zeolite/ Al2O3 1/16-in E
PNL/Grace CoMo/SiAl CoO/MoO3 3/13 13%Al2O3 SiO2 3/16-in T
Amoco NiMo/Y NiO/MoO3 3.5/18 Y-zeolite/ Al2O3 c 1/16-in E
BASF K8-11 CoO/MoO3 4.3/11 MgO spinel/ Al2O3 1 mm E
Akzo KF-742 CoO/MoO3 4.4/15.0 γ-Al2O3 1.3 mm Q
Akzo KF-840 NiO/MoO3 3.9/19.6 Al2O3c 1.3 mm Q
Criterion C-424 NiO/MoO3 4/19.5 Al2O3c 1.3 mm Q
Strem 78-166 Pt 5 γ-Al2O3 P
a E ) extrudates; R ) ring; T ) tablet; Q ) quadrilobe extrudates; P ) powder; and the size given is the o.d. b All three catalysts were used in a layered

bed. c Includes phosphorus oxide.

Table 2. Hydrotreating Hydrothermal Products Table 3. Hydrotreating Hydrothermal Products


Experimental Operating Conditions catalyst HT 400/S HT 400/S
catalyst CoMo 0402/S HT 400/S Ni-1404 feedstock TR7 TR12
feedstock distillate whole oil whole oil temperature, °C 398 397
temperature, °C 344 345 333 pressure, MPa 13.8 14.0
pressure, MPa 13.6 16.8 16.3 LHSV, vol oil/vol cat/h 0.10 0.11
oil feed rate, mL/h 342 204 274 H2 consumption, L/L oil 616 548
H2 feed rate, l/h 137 120 300 Yields
liquid hourly space velocity 0.38 0.24 0.30 total oil product, L/L feed 0.99 0.92
(LHSV), vol oil/(vol cat h) aqueous phase, L/L feed 0.20 0.20
H2 consumption, L/L oil 178 300 538 C5-225 °C, L/L feed >0.86 0.34
Experimental Results and Product Analyses gas (C1 to C4), wt % 14.1 9.0
carbon conversion, wt % carbon in aqueous 0.1 NA
to liquid product 70 77 69 Wet Product Analysis
to gas (C1 to C4) 1.0 2 15 H/C atomic ratio 1.65 1.5
to carbon on catalyst 12a 8 9 oxygen, wt % 0.0 0.8
Wet Product Analysis density, g/mL 0.84 0.91
H/C atomic ratio 1.72 1.75 1.82 C5-225 °C, vol % >87 37
oxygen, wt % 2.2 0.8 3.5-6.1
moisture content, wt % NA NA 1.7-2.9 while the TR12 contained more multiring phenolics. Under
density, g/mL 0.89 0.82 0.89 similar processing conditions as shown in Table 3, a lighter
sulfur in product, ppm NA 441 NA hydrocarbon product was produced from the TR7 oil. A need
a The distillate oil test was of limited duration, and the carbon on catalyst for additional hydrocracking function in the catalyst was
percentage was proportionally higher at 19%; but, the actual carbon weight indicated for use with TR12. The relationships of space velocity
percentage was similar to the other tests at 12%, suggesting that the carbon to deoxygenation and gasoline yield when using the HT400-S
laydown occurs early in the run and reaches a steady state.
catalyst are shown in Figures 1 and 2 for the two high-pressure
liquefaction bio-oils.16
similar to the sulfided CoMo catalyst except for having a much
Catalyst stability was also evaluated and neither sulfur loss
higher gas yield and requiring hydrogen consumption. The Ni
nor carbon buildup was judged to be the cause. The alkali
catalyst seemed to lose activity over the several hours of
content of the TR12 oil and its deposition on the catalyst was
operation as shown by the two product analyses in Table 2.
concluded to be the cause of catalyst deactivation by a pore
Distillations of these products showed that they were mostly
plugging mechanism over a 48 h test. The resulting loss of
gasoline range material (50-225 °C BP). GC-MS analysis
activity in the mixed Haldor Topsoe catalyst system is shown
showed that the cyclic ketones and phenolics in the liquefaction
in Figure 3.
product were converted to cyclic alkanes and aromatics.
Extensive studies with several catalysts provided a useful set
In further studies,15 comparisons were made between the of results for comparing residual oxygen content and hydrogen
whole oil presented above (TR7 water slurry, single pass) and to carbon ratio as well as gasoline yield as a function of space
an alternative form (TR12 produced with oil recycle). By GC- velocity16 (see Figures 4 and 5).
MS analysis of the two feed oils, it was found that the TR7 oil It was concluded that high yields of high-quality gasoline
contained primarily cyclic ketones and single-ring phenolics, can be produced from biomass-derived oils; however, low space

(15) Baker, E. G.; Elliott, D. C. Catalytic Upgrading of Biomass Pyrolysis (16) Baker, E. G.; Elliott, D. C. Catalytic Hydrotreating of Biomass-
Oils. In Research in Thermochemical Biomass ConVersion; Bridgwater, A. Derived Oils. In Pyrolysis Oils from Biomass: Producing, Analyzing and
V., Kuester, J. L., Eds.; Elsevier Science Publishers, LTD.: Barking, Upgrading; ACS Symposium Series 376; Soltes, E. J., Milne, T. A., Eds.;
England, 1988; pp 883-895. American Chemical Society: Washington, D.C., 1988; pp 228-240.
DeVelopments in Hydroprocessing Bio-oils Energy & Fuels, Vol. 21, No. 3, 2007 1795

Figure 1. Oxygen content of hydrotreated oils (400 °C, 13.8 MPa).

Figure 2. Yield of gasoline boiling range (C5-225 °C) oil (400 °C, 13.8 MPa).

velocities are required. Cracking and hydrogenation of the higher maximize the yield of aromatic gasoline by removing a frac-
molecular weight components are the rate-limiting steps in tion from the system before the rings became saturated.
upgrading biomass-derived oils. Future catalyst development Furthermore, hydrogen consumption was minimized by using
should be directed at these reactions. it only to hydrodeoxygenate the oxygen-containing compo-
An advanced process concept was also tested wherein the nents and it was not wasted on the saturation of aromatic
initial hydrotreating for oxygen reduction was followed by compounds.
separation of the aqueous, gasoline, and gas phases and a sec-
(17) Baker, E. G.; Elliott, D. C. Method of Upgrading Oils Containing
ond processing step of hydrocracking the heavy components.17 Hydroxyaromatic Hydrocarbon Compounds to Highly Aromatic Gasoline.
By performing the multistep process, it was possible to U.S. Patent Number 5,180,868, January 19, 1993.
1796 Energy & Fuels, Vol. 21, No. 3, 2007 Elliott

Figure 3. Catalyst deactivation with Haldor Topsoe catalyst and TR12 Oil (400 °C, 13.8 MPa).

Figure 4. Effect of catalysts on the quality of hydrotreated oil from TR7 (400 °C, 13.8 MPa).

In batchwise recycle tests, the results of the recycle of heavy test, as shown in Table 4. The two-stage hydrotreatment (with
components to a hydrotreating bed (recycle concept) or to a intermediate separation) results in a 13% reduction in hydrogen
hydrocracking environment (multistep concept) were compared. consumption for equivalent gasoline yield. The space velocity
Both concepts showed good gasoline yield, but the hydrocrack- is increased by a factor of 4. With the use of a hydrocracking
ing catalysts were more active toward molecular weight reduc- step, there was a similar reduction in hydrogen consumption
tion and produced more gasoline. These batchwise recycle tests and a similar fourfold increase in the space velocity resulting
were even more impressive when compared to the single-pass in a gasoline yield increase of up to 30%.
DeVelopments in Hydroprocessing Bio-oils Energy & Fuels, Vol. 21, No. 3, 2007 1797

Figure 5. Effect of catalysts on the quality of hydrotreated oil from TR12 (400 °C, 13.8 MPa).

Table 4. Comparison of Two-Step and Single-Step Resultsa


two-stage single-stage
CoMo 499 (CoMo-HT in 2nd) KAT 4000 (CoMo-HC in 2nd) CoMo HT only
H2 consumed, L/L 500 503 575
LHSV, L/(L h) 0.19 0.19 0.05
gasoline yield, L/L 0.62 0.80 0.61
temperature, °C 400/400 400/435 400
pressure, MPa 13.8 13.8 13.8
a TR12 bio-oil feed was used in the HT step, distilled to remove volatiles <160 °C before the 2nd step.

In comparison with the biomass-derived oils, the hydrotreated More detailed analysis of several gasoline-range distillates
products are significantly upgraded. The oxygen content is from the hydrotreated biomass-derived oils was undertaken.19
greatly reduced and, coincidentally, so is the density of the Those analyses provided additional detail on the makeup of the
products. The density difference has a significant impact products and also further substantiated the relationships of the
because, although the mass yield of the hydrotreated products product composition to product properties. As seen in Table 5,
is in the range of 80%, the volume yield in many cases exceeds elemental composition was compared with component fraction-
100%. A primary concern throughout the research at PNNL was ation and component analysis by instrumental methods. To
the maintenance of the aromatic character of the biomass oil in fractionate the components of the distillates, the ASTM D-1319
order to minimize hydrogen consumption and to produce a method was used for determining hydrocarbon types by
higher octane gasoline blending stock, as recorded in one of fluorescent indicator adsorption. By nuclear magnetic resonance
their patents.18 The extent of saturation as shown by the H/C (NMR) of carbon-13, similar component groups were identified
ratio is a useful indicator of the aromatic character of the and quantified.
product. Saturation of the aromatic components has a strongly For most of the samples listed in Table 5, the D-1319 data
deleterious effect on the octane of the product. A review of the compared quite favorably with the C-13 NMR results. The
literature shows that cyclic hydrocarbons have poor octanes aromatic and aliphatic portions were nearly identical. The
similar to straight-chained hydrocarbons. Although the crude D-1319 method consistently showed a small olefin fraction in
hydrotreated products did contain minor amounts of oxygen, the oil, while the NMR analysis detected essentially no olefinic
water solubility in the products was low. In addition, although carbon atoms. Further analysis of the fractions from the D-1319
sulfided catalysts were used in the hydrotreating, little incor- separation was performed by gas chromatography with a mass
poration of sulfur into the nearly sulfur-free biomass oils selective detector (HP 5970). Individual components in each
occurred. fraction were identified and semiquantitatively determined by

(18) Baker, E. G.; Elliott, D. C. Method of Upgrading Oils Containing (19) Elliott, D.C.; Schiefelbein, G. F. Liquid Hydrocarbon Fuels from
Hydroxyaromatic Hydrocarbon Compounds to Highly Aromatic Gasoline. Biomass. Prepr. Pap.sAm. Chem. Soc., DiV. Fuel Chem. 1989, 34 (4),
U.S. Patent Number 5,180,868, January 19, 1993. 1160-6.
1798 Energy & Fuels, Vol. 21, No. 3, 2007 Elliott

Table 5. Distillate Products from Hydrotreatment


aromatic/
elemental aliphatic/ octane
analysis, % 13C NMR olefin numbers
density HHV gasoline boiling aromatic/ actuala
C H H/C ratio O (g/mL) (Btu/lb) (IBP 225 °C) range (°C) aliphatic aromatic D-1319 MON RON R+M/2
86.6 12.1 1.66 1.3 0.844 100% 23-225 28/72 43% 44.1/55.1/0.8 72.0 77.0 74.5
85.4 12.5 1.74 2.2 0.791 100% 68-176 29/71 25.4% 39.5/53.6/6.9
87.1 12.0 1.64 0.9 0.859 100% 23-225 30/70 29.0% 47.4/48.3/4.3
86.2 13.1 1.81 0.6 0.823 18990 100% 23-225 24/76 32% 33.9/63.3/2.8 72.8 78.1 75.5
0.810 100% 23-165 20/80 28% 28.3/69.9/1.8
86.0 12.7 1.75 1.3 0.803 100% 72-157 22/78 29% 33.7/59.1/7.2
84.3 13.7 1.93 1.5 0.782 100% 57-183 12/88 18% 18.1/77.1/4.8
85.6 13.3 1.84 1.2 0.802 100% 63-149 16/84 20% 28.3/68.6/3.1
a “Actual aromatic” is the sum of the aromatic carbon and the nonaromatic substituents on the aromatic rings.

Table 6. Components of D-1319 Chromatography Fractions (within Table 7. Hydrotreating Hydrothermal Products
Each Fraction, from Highest Total Ion Count)
desalted isooctane xylene
saturated hydrocarbons olefinic hydrocarbons only extract extract
ethylcyclohexane octahydroindene Feed Oil Composition, as Fed
propylcyclohexane octahydropentalene carbon, wt % 70.6 80.6 81.0
methylethylcyclohexane methyloctahydropentalene hydrogen, wt % 8.2 9.4 8.2
methylcyclohexane oxygen, wt % 15.5 8.2 9.5
methylpropylcyclohexane H/C ratio 1.38 1.39 1.20
methylethylcyclohexane temperature, °C 350 350 350
methylpropylcyclopentane pressure, MPa 10 10 10
ethylpropylcyclohexane LHSV, vol oil/(vol cat h) 1.2 0.79 1.13
dimethylcyclohexane H2 feed, L/L oil 472 773 500
methylcyclopentane H2 consumption, L/L oil NA 81 152
Product Analysis
aromatic hydrocarbons alcohol soluble components
H/C atomic ratio 1.44 1.38 1.35
ethylmethylbenzene dimethylphenol oxygen, wt % 2.3 4.8 2.5
methylpropylbenzene naphthalene coke on catalyst, wt % 11.0 21.4 11.1
propylbenzene ethylphenol C5-225 °C, wt % (naphtha) 20.0 13.3 17.9
C4-alkyl-benzene cresol C13-C20, wt % (atm gas oil) 36.5 45.9 37.7
C2-alkyl-tetralin ethylmethylphenol C21-C44, wt % (vac gas oil) 29.8 33.5 29.7
methyltetralin cresol residue 13.6 7.3 14.5
tetralin methylnaphthalene
methylindan ethylmethylphenol varied from 50 to 150 bar and from 300 to 390 °C for 20 min
C5-alkyl-benzene dimethylphenol to 5 h. Typical yields at 350 °C and 100 bar were 1% gas, 16%
methylpropylbenzene ethyl phenol naphtha, 41% atmospheric gas oil, 34% vacuum gas oil, and
8% vacuum residue from a feedstock which contained 2.5%
the intensity of the total ion current for each peak. Components naphtha, 38.2% atmospheric gas oil, 39.5% vacuum gas oil, and
in each of the fractions were listed in Table 6. With this analysis, 19.8% vacuum residue (based on gas chromatography simulated
the NMR results were confirmed, as the primary components distillation). Coke laydown on the catalyst was evaluated and
of the olefin fraction were found to be bicyclic components. found to be at a minimum between 350 and 375 °C. The coke
Some difficulty was encountered with this analysis because of deposition amounted to 2-4 wt % on the catalyst, appeared to
the small fraction size and the contamination by the aliphatic occur during the first 2.5 h of the test, and did not increase
fraction. However, no mass spectra of olefin components were with longer time at the same temperature. The operating
confirmed, and the primary components in the fraction could temperature was identified as the most important factor affecting
be determined by comparison with the aliphatic fraction analysis. yields with the pressure effect being much less. At higher
Chalmers Institute of Technology. Chalmers also performed temperature, more of the heavier components in the feed were
catalytic hydroprocessing research using the TR12 product oil converted to lighter components. By extrapolation, it was found
from the U.S. high-pressure (hydrothermal) biomass liquefaction that, if the reaction could be done at 410 °C, then the residue
effort at Albany, Oregon. The effort focused on preprocessing would be eliminated.
of the oil to facilitate subsequent catalytic processing. Solvent In subsequent tests, the wood oil was desalted (water washed
extraction was evaluated as a means to separate a light oil from while in isooctane solvent) before hydroprocessing in a down-
the heavier components including the mineral material (residual flow, continuous-feed, fixed-bed reactor using the same cata-
sodium salts derived from the liquefaction catalyst). Extraction lyst.22 The three feedstocks listed in Table 7 are all desalted;
yields varied from 74% with acetone to 56% with xylene, to the last two were sequentially extracted with the indicated
30% with octane or decalin, and only 3% with pentane.20 solvent from the desalted oil. In tests performed at 350 °C and
The decalin extract of the TR12 oil was used as a feedstock 10 MPa of pressure, 100 ppm carbon disulfide was added to
to batch hydroprocessing tests using the Akzo Ketjen 742 maintain activity of the Ketjen 742 catalyst. The nonextracted
catalyst (sulfided CoMo on alumina).21 Conditions evaluated but desalted oil was more readily hydrotreated than the extracted
oils, and the authors concluded that desalting was sufficient with
(20) Gevert, B. S.; Otterstedt, J.-E. Upgrading of Liquefied Biomass to
Transportation Fuels by Extraction. Energy from Biomass & Wastes X; no need for extraction before hydroprocessing.
IGT: Chicago, 1986; pp 845-854.
(21) Gevert, B. S.; Otterstedt, J.-E. Upgrading of Directly Liquefied (22) Gevert, B. S.; Andersson, P.; Jaras, S.; Sandqvist, S. Hydropro-
Biomass to Transportation FuelssHydroprocessing. Biomass 1987, 13, cessing of Desalted Directly Liquefied Biomass. Prepr. Pap.sAm. Chem.
105-115. Soc., DiV. Fuel Chem. 1988, 33 (4), 913-919.
DeVelopments in Hydroprocessing Bio-oils Energy & Fuels, Vol. 21, No. 3, 2007 1799

In subsequent testing, the use of wide-pore hydroprocessing (oxygen-containing components in the oil polymerized in the
catalysts showed no improvement for hydrodeoxygenation, reactor) and was abandoned.
hydrogenation, or hydrocracking of the desalted oil as shown No detailed product analysis is reported for these tests. Since
by product oxygen content, H/C ratio of the product, and the the operations were performed using a carrier solvent, actual
amount of distillation residue in the product, respectively.23 product recovery would have been a tedious process. With the
However, by pore volume analysis, it was determined that the diluted feedstock, any problems with coking of the pyrolysis
more active small-pore catalyst more readily lost pore volume oil would have been masked. The length of the tests is not
during use and would likely be less active over time, even reported, nor is any information provided on the stability of
though the carbon deposition measured by weight percent was the catalyst activity.
less.
Fast Pyrolysis Bio-oil Upgrading
Slow Pyrolysis Oil Upgrading
Upgrading of fast pyrolysis bio-oil is a difficult proposition
Some work in the 1970s envisioned the use of slow pyrolysis due to the reactivity of the condensed bio-oil recovered from
reactor systems for wood tar production as a biomass-derived fast pyrolysis. The hydrodeoxygenation tactic used with the
crude oil. Compared to fast pyrolysis, the slow pyrolysis oil high-pressure liquefaction (hydrothermal) bio-oil is still an
was recovered in lower mass yields but with lower oxygen and overriding concern in order to make a more useful product from
water content and better thermal stability. Initial research on the bio-oil. However, direct application of hydrotreating tech-
upgrading this oil had early success as a result of the more easily nology from petroleum processing is not possible with fast
processed oil. pyrolysis bio-oil.
Texas A&M University. Pyrolytic oil produced in a Tech- Pacific Northwest National Laboratory (PNL/PNNL). Bio-
Air reactor system from southern pine was the feedstock for oil upgrading work at PNNL has focused on catalytic hydro-
this research along with other tars produced in an updraft processing. The initial work as described above focused on the
gasifier. Batch and continuous-flow reactors were used, and 20 product from high-pressure liquefaction, a hydrothermal, alkali-
catalyst formulations were tested.24 Decahydronaphthalene was catalyzed process which produced a more deoxygenated product
used as a solvent carrier at a nominal 2-to-1 ratio with the than fast pyrolysis bio-oil. As a result, the hydrothermal product
pyrolysis oil in all these tests. On the basis of batch reactor test was more thermally stable and required less hydrogenation to
results at 400 °C for 1 h, the Pd on alumina catalyst was produce a gasoline product. An initial test with fast pyrolysis
determined to be most useful with the highest liquid yield. Pt bio-oil clearly demonstrated that the hydrotreating approach for
or Re on alumina and Raney nickel were nearly as useful but hydrothermal product was inappropriate for bio-oil26 as it
produced higher gas yields. Ru and Rh were found to be the resulted in heavy product tar plugging the reactor system and
most active for gas formation. The use of a carbon support in the catalyst bed encased in a cokelike product. The bio-oil from
place of the alumina support caused more gas production. a poplar wood pyrolyszed in a fluid bed reactor was hydrotreated
Sulfided CoMo, NiMo, and NiW catalysts were found to have over a sulfided CoMo catalyst at 355 °C and 13.8 MPa with a
much lower activity compared to the precious metal catalysts. liquid hourly space velocity (LHSV) of 0.35 and a hydrogen
Subsequent work focused on processing the Tech-Air pine flow at 318 std L/L of bio-oil. Hydrogen consumption was
pyrolysis oil using the 5% Pt on alumina catalyst and com- calculated at 127 L/L. There was only a 23% mass yield of a
mercial CoMo, NiMo, and NiW catalysts in the continuous- liquid product which had 3.6 and 5.9% oxygen with H/C atomic
flow reactor.25 The reactor system was a trickle-bed type with ratios of 1.55-1.45.
a 19 mm i.d. and 813 mm long, of which only the top 508 mm Two-Stage Hydrotreatment. In order to address the instability
were packed with catalyst with inert 3 mm Pyrex beads in the of the bio-oil under hydrotreating conditions, a two-stage process
bottom. The temperature of the reactor bed ranged from 190 was developed. Catalytic hydrotreatment at temperatures below
°C up to the reaction temperature, which varied from 350 to 300 °C with either Ni or sulfided CoMo catalyst were effective
400 °C. Hydrogen was fed at 107 vol/vol of oil, and the reactor in producing a stabilized oil product. This concept was patented
pressure was maintained at from 5.3 to 10.4 MPa. The oil was by Battelle at the time.27 In the attempt at a higher temperature
processed at a weight hourly space velocity (WHSV) from 0.5 (310 °C), the reactor bed plugged with solid cokelike material.
to 3. As reported,28 catalyzed hydrotreatment produced a viscous,
The Pt catalyst (Strem 78-166 mixed with 20% silica and black oil with a density of about 1, containing about 25% oxygen
extruded) was found to be the most active for oxygen removal. and an H/C atomic ratio of about 1.5. The liquid yield was about
Oxygen removal ranged from 27 to 45% over the temperature 0.42 L/L from a hardwood bio-oil produced in an entrained-
range, while that of NiMo and CoMo ranged from 15 to 39%. flow reactor. The presence of a hydrogenation catalyst was
At the highest operating pressures, the oxygen removal was critical as shown by the result in the test where only an alumina
increased to 51%. Product alkanes and aromatics increased to spacer was present wherein the bed plugged with cokelike
almost 20% using the Pt catalyst at the highest temperature at material. Processing data are given in Table 8.
a WHSV of 2 and 8.7 MPa, while they never exceeded 13% Further testing of the low-temperature (stage one) hydropro-
using the CoMo catalyst and this was only 8% with NiMo. The cessing demonstrated that the space velocity and hydrogen flows
NiW was reported to not be effective in oxygen removal
(26) Elliott, D. C.; Baker, E. G. Biomass Liquefaction Product Analysis
(23) Gevert, S. B.; Andersson, P. B. W.; Sandqvist, S. P. Hydroprocessing and Upgrading. Comptes Rendus de l’Atelier de TraVail sur la Liquéfaction
of Directly Liquefied Biomass with Large-Pore Catalysts. Energy Fuels de la Biomasse; report 23130, NRCC: Sherbrooke, Quebec, Canada,
1990, 4, 78-81. September 29-30, 1983; pp 176-183.
(24) Soltes, E. J.; Lin, S.-C. K.; Sheu, Y.-H. E. Catalyst Specificities in (27) Elliott, D. C.; Baker, E. G. Process For Upgrading Biomass
High Pressure Hydroprocessing of Pyrolysis and Gasification Tars. Prepr. Pyrolyzates. U.S. Patent Number 4,795,841, January 3, 1989.
Pap.sAm. Chem. Soc., DiV. Fuel Chem. 1987, 32 (2), 229-239. (28) Elliott, D. C.; Baker, E. G. Catalytic Hydrotreating of Biomass
(25) Sheu, Y.-H. E.; Anthony, R. G.; Soltes, E. J. Kinetic Studies of Liquefaction Products to Produce Hydrocarbon Fuels: Interim Report;
Upgrading Pine Pyrolysis Oil by Hydrotreatment. Fuel Process. Technol. report no. PNL-5844, Pacific Northwest National Laboratory: Richland
1988, 19, 31-50. Washington, 1986.
1800 Energy & Fuels, Vol. 21, No. 3, 2007 Elliott

Table 8. Stage One Hydroprocessing of Bio-oil oil. The results are presented in Table 11.15 RTP indicates the
Experimental Operating Conditions Ensyn, rapid thermal processing; AFP indicates atmospheric
catalyst Ni3266 Ni3266 Ni3266 HT-400 R-Al2O3 flash pyrolysis, such that ablative is the SERI (now NREL, the
temperature, °C 258 280 310 273 254 National Renewable Energy Laboratory) vortex reactor and
pressure, MPa 14.0 14.2 14.2 14.0 13.8
oil feed rate, mL/h 290 396 405 392 411
fluid-bed is the University of Waterloo product. In the noniso-
H2 feed rate, L/h 168 216 240 168 180 thermal processing, the gas yield is higher but the carbon loss
LHSV, vol oil/ 0.32 0.44 0.45 0.44 0.46 to the aqueous phase is lower. Processing the aqueous phase
(vol cat h) through the high-temperature hydrotreating process apparently
H2 consumption, 66 161 252 135 -49 converts the aqueous phase carbon to gas with an additional
L/L oil
consumption of hydrogen gas.
Experimental Results and Product Analyses On the basis of material and elemental balance calculations,
carbon conversion, wt % these results were produced only at near-steady state. A
to oil/aqueous 36/14 57/10 0/5 55/11 ND
to gas (C1 to C4) 9 11 16 9 10 calculation shows that steady-state operation would be expected
to carbon on catalyst 9 8 ND ND ND to produce 0.50-0.55 L of product oil/L of feed. The product
oil product yield, 0.28 0.42 reactor 0.42 reactor oil would be expected to contain 2-3 wt % oxygen and would
L/L feed plugged plugged have an H/C ratio of 1.5 and a density of 0.92 g/mL. About
Wet Product Analysis 50-60% of the oil would be in the gasoline boiling range; of
H/C atomic ratio 1.54 1.42 1.49/1.63 1.47 1.14 the remainder, about 30% would be distillable gas oil (diesel)
oxygen, wt % 26.8 25.0 19.4/13.2 24.6 24.9
density, g/mL 1.1 ND ND/0.96 ND ND
and the rest would be resid. Gasoline yield and residual oxygen
content in the product can both be directly related to the
both played a minor role in the processing results. As seen in processing space velocity.15
Table 9, space velocity variation in the range of 0.44-0.62 (three With the peat oil,31 the results in Table 11 represent steady-
leftmost columns) has an effect on the extent of oxygen removal state operation. The quality of the product oil from peat is
and hydrogen consumption, while at higher space velocities up somewhat better than the hydrotreated wood-derived oils. The
to 1.6 the results remain about level. Reduction of the hydrogen peat product contained more paraffinic material (including
flow to zero at moderate space velocity (three rightmost alkylcyclohexanes), which accounts for the higher H/C ratio.
columns) reduced the hydrogen incorporation noticeably without However, it also contained 1.1 wt % nitrogen, a level that is
affecting oxygen removal. The product oil quality was reduced seldom seen with wood-derived products.
as measured by density and viscosity. More extensive testing was done with the bio-oil produced
The thermal stability of the low-temperature, hydrotreated by vacuum pyrolysis at the University of Laval in Canada. The
product oil (it could be distilled batchwise without coke results32 from nonisothermal catalytic hydrotreating of primary
formation in the pot) as well as its elemental composition (70.7% condenser oil are presented in Table 12. The first test was
C, 8.1% H, 0.1% N, 20.9% O, calculated to a dry basis) performed at relatively low temperatures and a high bio-oil feed
indicated that it was significantly upgraded from the original rate. As a result, the product quality is low, as shown by high
bio-oil. Chemical composition analysis by GC-MS verified that oxygen content and density. The second test was performed at
the carbonyl components were destroyed as well as the conditions optimized to improve product qualityshigher tem-
saturation of the olefinic side chains. The relative amount of perature and lower feed rate. Hydrogen consumption was higher
phenolic components increased at the expense of the aromatic as was gas formation. Because of the reduced density of the
ethers. Saturated cyclic alcohols were also present indicating product, the volumetric yield increased to over 0.4 L/L based
some hydrogenation of the aromatic rings. on bio-oil feed. This data set probably does not represent steady-
On the basis of these results, a two-stage process for state operation. The last two data sets are derived from longer-
hydrotreating fast pyrolysis bio-oil to liquid fuels was demon- term tests (18-22 h). They are believed to better represent
strated at bench scale. Gasoline-range hydrocarbons were steady-state operation after break-in of the catalyst.
produced from low-temperature hydrotreated bio-oil, using a These process tests can be further compared by plotting the
sulfided cobalt molybdenum catalyst, at conditions of ap- product quality versus processing rate.33 In Figure 6, oxygen
proximately 350 °C and 13.8 MPa. Carbon conversion to content in the upgraded product oil is plotted as a function of
gasoline was in excess of 80 wt % with a liquid product yield space velocity for nonisothermal catalytic hydrotreating of bio-
of 0.77 L/L. Hydrogen consumption was measured at 728 L/L oils. In this plot, the data for all the atmospheric fast pyrolysis
of oil feed based on gas-phase measurements in and out of the processes (flash pyrol.) appear to fall on the same line. We
reactor. The space velocity in the upflow reactor was relatively conclude that the oil source components in all the feedstocks
low at 0.07 vol of oil/(vol of catalyst h).29 The additional results are essentially the same, so that the same type of product is
of the two-stage processing are presented in Table 10, as well produced from all the tests. However, the vacuum pyrolysis oil
as calculation of the combination of the two steps.30 (vac. pyrol.) appears to be somewhat more difficult to deoxy-
Nonisothermal Hydrotreatment. To simplify the process and genate to low levels. In Figure 7, gasoline yield is plotted versus
maximize liquid product yield, the two stages were combined space velocity. In this comparison, the atmospheric fast pyrolysis
in a single nonisothermal catalyst bed. This concept was tested upgraded products again fall on the same line, suggesting an
at the bench-scale, in an upflow configuration, with several
(31) Elliott, D. C.; Baker, E. G.; Piskorz, J.; Scott, D. S.; Solantausta,
wood-derived bio-oil feedstocks and also a peat fast pyrolysis Y. Production of Liquid Hydrocarbon Fuels from Peat. Energy Fuels 1988,
2, 234-235.
(29) Elliott, D. C.; Baker, E. G. Hydrotreating Biomass Liquids to (32) Elliott, D. C. Perform Hydrotreatment of Biomass Liquefaction
Produce Hydrocarbon Fuels. Energy from Biomass & Wastes X; IGT: Products Including Vacuum Pyrolysis Oils; contract no. 2312012867, letter
Chicago, 1986; pp 765-784. report to Zeton, Incorporated: Burlington, Ontario, October 20, 1988.
(30) Baker, E. G.; Elliott, D. C. Catalytic Upgrading of Biomass Pyrolysis (33) Elliott, D. C. Perform Hydrotreatment of Biomass Liquefaction
Oils. Research in Thermochemical Biomass ConVersion; Bridgwater, A. Products RTP (Rapid Thermal Processing) LiquidsEnsyn Engineering
V., Kuester, J. L., Eds.; Elsevier Science Publishers, LTD.: Barking, Associates, Inc.; contract no. 2312012867, letter report to Zeton, Incorpo-
England, 1988; pp 883-895. rated: Burlington, Ontario, February 8, 1989.
DeVelopments in Hydroprocessing Bio-oils Energy & Fuels, Vol. 21, No. 3, 2007 1801

Table 9. Effect of Space Velocity and Hydrogen Flow in Stage One Hydroprocessing
Experimental Operating Conditions
catalyst HT-400
temperature, °C 273 271 271 274 271 270 277 276
pressure, MPa 14.0 14.0 14.0 13.9 14.0 14.1 14.0 13.9
oil feed rate, mL/h 392 515 555 935 1200 1440 934 959
H2 feed rate, L/h 168 120 120 120 120 120 40 0
LHSV, vol oil/ (vol cat h) 0.44 0.57 0.62 1.04 1.33 1.60 1.04 1.07
H2 consumption, L/L oil 135 90 60 39 32 28 26 0
Experimental Results and Product Analyses
carbon conversion, wt %
to oil/aqueous 55/11 80/10 87/8 83/5 83/7 87/11 83/9 85/7
to gas (C1 to C4) 9 7 7 5 4 4 10 10
oil product yield, L/L feed 0.42 0.56 0.69 0.66 0.65 0.70 0.61 0.61
Wet Product Analysis
H/C atomic ratio 1.47 1.47 1.56 1.56 1.48 1.58 1.44 1.33
viscosity, cPs @ 60 °C ND ND ND ND ND 14 200 18 700 32 700
density, g/mL @ 20 °C ND ND ND ND ND 1.14 1.16 1.18
oxygen, wt % 24.6 30.8 32.7 32.7 31.4 34.2 30.7 29.5
oxygen rejection ND 70 57 59 62 55 61 62

Table 10. Two-Stage Hydroprocessing of Bio-oil with oxygen contents less than 15% were low viscosity, as good
Experimental Operating Conditions or better than the feed oil. The products with 15 to 25% oxygen
feed oil bio-oil stage 1 product combined were several orders of magnitude higher in viscosity and difficult
catalyst HT-400 HT-400 HT-400 to recover from the reactor system. In the upflow mode, there
temperature, °C 274 353
pressure, MPa 14.1 14.2
is a tendency for the low-viscosity (and also high-volatility)
LHSV, vol oil/(vol cat h) 0.62 0.11 products to fractionate from the more viscous, low-volatility
products and leave the reactor with the excess hydrogen. In this
Experimental Results and Product Analyses manner, the catalyst bed slowly fills with high-viscosity oil
H2 consumption, L/L oil 60 576 457
carbon conversion, wt % which exits the reactor in disproportionately high yields later
to aqueous 8.2 0.7 8.8 in the experiment.
to gas (C1 to C4) 7.0 12.6 17 The products from the near-steady-state operations appear
oil product yield, L/L feed 0.69 0.62 0.43
aqueous phase yield L/L feed 0.34 0.38 0.61
much heavier than expected as a result of the fractionation
C5-225 °C, L/L feed 0.07 0.45 0.31 occurring during the initial period of the experiments. A
Product Inspections
significant period of non-steady-state operation occurs at the
H/C atomic ratio 1.58 1.67 beginning of the tests in which product recovery is incomplete
oxygen, wt % 32.7 2.3 while volatile products are swept from the reactor and heavier
density, g/mL 1.14 0.86 (liquid-phase) products slowly build up in the catalyst bed.
C5-225 °C, vol % 10 72 Eventually (after several hours online), the catalyst bed fills with
viscosity, cPs @ 60 °C 14,200 ND
moisture, wt % 14.1 ND liquid phase, the “whole” product oil begins to exit the reactor,
and a near-steady-state operation is achieved. However, there
equivalent yield, while the vacuum pyrolysis bio-oil gives a remains the accounting of the light products recovered in the
somewhat higher yield of the gasoline-range product by early stages of the test. These must be added back to the heavy
hydrotreating. whole products to account for a true mass balance. These initial-
Low-SeVerity Hydrotreatment. Using the above results, a period products are the type of light fuel oil which is being
prediction of results under less-severe processing conditions was sought after for the desired turbine-fuel product.
determined.34 As shown in Figure 8, an extrapolation of data In these tests wherein the vapor-phase product and the liquid-
to higher space velocity processing can be used to predict the phase product are so different, it is difficult to bring short-term
oxygen content of hydrotreated biomass fast pyrolysis oils. tests to steady-state operation. The fixed-bed reactor in an
Using this data, it was predicted that a space velocity of 0.4 upflow mode of operation naturally allows product fractionation.
would produce a product with an oxygen content of ap- In addition, the difference in product properties makes control
proximately 15%. Such an oil should require less hydrogen of the process equipment and recovery of the product more
consumption and would be produced in higher yield because difficult than in gasoline-production tests with a totally vapor-
there would be less gas formation. This prediction was based phase product.
on continued use of the high-temperature regime (up to 400 Catalyst analysis after these initial low-severity tests indicated
°C) and 13.8 MPa of pressure in the reactor. some changes had occurred. There was clear evidence in some
Initial low-severity hydrotreating tests were made using an of the catalyst pellets of the reaction of the alumina support to
upflow configuration.35 It is difficult to operate in the upflow the hydroxide (boehmite). This reaction in aqueous processing
mode because of the tendency of the product to fractionate and under similar conditions has been reported.36,37 The high-surface-
plugging in the reactor by high-viscosity products. The products area aluminas (gamma and delta) are not stable in the presence
of water at 350 °C. The stability of the sulfide phases was not
(34) Elliott, D. C. Biomass Fast Pyrolysis OilsLow-SeVerity Hydrotreat-
ing; contract no. AIR-CT-92-0216, final report to Arusia: Perugia, Italy,
December 1996. (36) Elliott, D. C.; Sealock, L. J., Jr.; Baker, E. G. Chemical Processing
(35) Elliott, D. C.; Hart, T. R.; Neuenschwander, G. G.; McKinney, M. in High-Pressure Aqueous Environments. 2. Development of Catalysts for
D.; Norton, M. V.; Abrams, C. W. EnVironmental Impacts of Thermo- Gasification. Ind. Eng. Chem. Res. 1993, 32, 1542-1548.
chemical Biomass ConVersion; NREL/TP-433-7867, National Renewable (37) Laurent, E. Etude et contrôle des réactions d’hydrodésoxygénation
Energy Laboratory: Golden, Colorado, 1995 (also PNL-10413, Pacific lors de l’hydroraffinanage des huiles de pyrolyze de la biomasse. Ph.D. Thesis.
Northwest National Laboratory: Richland, Washington). Université Catholique de Louvain, Louvain-la-Neuve, Belgium, 1993.
1802 Energy & Fuels, Vol. 21, No. 3, 2007 Elliott

Table 11. Nonisothermal Hydroprocessing of Bio-oil


Experimental Operating Conditions
feed oil circulating bed RTP (poplar) ablative pyrol AFP (pine) fluid-bed AFP (peat)
catalyst CoMo479/499 HT-400 HT-400
temperature, bottom/top, °C 252/392 259/376 302/391
pressure, MPa 13.8 13.7 13.8
LHSV, vol oil/(vol cat h) 0.12 0.10 0.19
Experimental Results and Product Analyses
H2 consumption, L/L oil 588 689 498
carbon conversion, wt %
to aqueous 0.4 0.6 1.0
to gas (C1 to C4) 27.2 25.0 30.0
oil product yield, L/L feed 0.30 0.37 0.44
Aqueous phase yield l/l feed NA 0.51 0.34
C5-225 °C, L/L feed 0.26 0.27 0.36
Product Inspections
H/C atomic ratio 1.65 1.68 1.80
oxygen, wt % 1.7 1.3 1.5
density, g/mL 0.86 0.85 0.86
C5-225 °C, vol % 81.1 73 82
Gasoline Inspections
H/C atomic ratio 1.64 NA NA
oxygen, wt % 0.9 NA NA
density, g/mL 0.859 NA NA
saturates, % 48.3 NA NA
aromatics, % 47.4 NA NA
olefins, % 4.3 NA NA

Table 12. Nonisothermal Hydroprocessing of Vacuum Pyrolysis configuration compared to earlier upflow tests. High levels of
Bio-oil hydrogen consumption and high product quality were seen even
Experimental Operating Conditions at much higher space velocity. Product oils were analyzed to
feed oil hardwood hardwood softwood softwood quantify the effect of space velocity on the product properties.
catalyst CoMo 479/499
temperature, 262/371 258/400 266/386 252/382
The two catalysts had significantly different activities. Without
bottom/top, °C a strong hydrotreating catalyst effect in the reactor, exothermic
pressure, MPa 14.0 14.1 14.0 14.1 pyrolysis condensation reactions caused major temperature
LHSV, 0.28 0.13 0.11 0.13 excursions and led to coke buildup on the catalyst. Buildup of
vol oil/(vol cat h) heavy tar products in the reactor and effluent lines caused
Experimental Results and Product Analyses plugging and premature experiment terminations. Product
H2 consumption, 379 711 613 727 fractionation occurred, similarly to the upflow experiments, in
L/L oil
carbon conversion 23.9 35.5 33.9 18.4 which the excess hydrogen gas exiting the reactor carried a
to gas, wt % substantial portion of the light products out while leaving most
oil product yield, 0.29 0.42 0.51 0.40 of the biocrude still as a heavy tar product. Total carbon balances
L/L feed were difficult to measure, but very low space velocities would
C5-225 °C, L/L feed NA 0.37 0.41 0.34
be required to effectively hydrotreat the biocrude over this
Product Inspections catalyst.
H/C atomic ratio 1.47 1.73 1.53 1.67
oxygen, wt % 9.0 0.8 2.9 3.6 The viscosity of the product oil is also a function of the
density, g/mL 0.93 0.83 0.87 0.865 oxygen content, as seen in Figure 9. The range shown goes all
C5-225 °C, vol % NA 76 73 76 the way from the low viscosity required for turbine fuels of
quantified, but there was no nickel or molybdenum oxide evident less than 5 cps to the heavy tar products with high oxygen
by X-ray diffraction while the sulfides were both identified. contents and viscosities >100 000 cps whose pour points would
The results presented below were produced in the continuous be around room temperature. These results suggest that only
feed catalytic hydrotreater operated in a downflow configura- the highly upgraded oils with oxygen contents of 5% or less
tion.38 The tests were made with a NiMo on alumina (Haldor have the potential for direct use as turbine fuels because of
Topsoe TK751, 1 mm extrudates) conventional hydrotreating viscosity limitations. Figure 9 also includes the data for the three
catalyst, presulfided processing an ablative reactor bio-oil gener- bio-oils. The raw bio-oils show decreasing viscosity with
ated from poplar hardwood by NREL. A second catalyst was increasing oxygen content because of increasing water content
also tested (BASF K8-11). This CoMo catalyst is unlike the in the raw biocrude.
NiMo in that the support is a spinel not the conventional high- Figure 10 shows the effect of temperature on the viscosity
surface-area γ-alumina commonly used in petroleum hydrotreat- of several of the heavy oil products. These results suggest that
ing catalysts. Better chemical stability of the support was the highly oxygenated products could not be used as turbine
envisioned; but, the catalyst, which is not normally used for fuels because of their high viscosity even with preheating. The
petroleum hydrotreating, exhibited much lower activity overall. 10% oxygen oil might be used as turbine fuel with preheating
The results of these tests are summarized in Table 13 below. to 50 °C or higher.
Higher levels of conversion were seen in the downflow Organic contamination of the water byproduct from hy-
drotreating is a concern relative to the overall wastewater
(38) Elliott, D. C.; Neuenschwander, G. G. Liquid Fuels by Low-Severity treatment requirements for the plant. Figure 11 clearly shows
Hydrotreating of Biocrude. DeVelopments in Thermochemical Biomass
ConVersion; Bridgwater, A. V., Boocock, D. G. B., Eds.; Blackie Academic that the contamination of the byproduct water as represented
& Professional: London, 1996; Vol. 1, pp 611-621. by the carbon content increases with the oxygen content of the
DeVelopments in Hydroprocessing Bio-oils Energy & Fuels, Vol. 21, No. 3, 2007 1803

Figure 6. Effect of space velocity on product oxygen content.

Figure 7. Effect of space velocity on the yield of gasoline range hydrocarbon.

lyptus oil from fluid-bed pyrolysis (Union Fenosa). Table 14


provides some of the results. Although the eucalyptus-oil test
was performed at higher temperature and lower space velocity,
the product-oil quality was significantly lower, with higher
oxygen content and density. Further substantiating the conclu-
sion of lower reactivity is the lower gas yield and also the lower
hydrogen consumption. The combination of high viscosity and
poor pumping performance with the lower reactivity make the
eucalyptus oil more difficult to hydrotreat.
A number of comparisons were made between the PNNL
work and that published by Veba Oel.39 The experimental
reactor results are evaluated in terms of deoxygenation which
Figure 8. Processing rate effect on product oxygen content. incorporates the effects of temperature, pressure, and catalyst.
The deoxygenation is defined as the percent of chemically
product oil up to a range of 5-10 wt % when the product oil
combined oxygen in the bio-oil removed when compared to
oxygen content is >10%. At that level, the oil-water separation
the oil product. Deoxygenation as a function of space velocity
is difficult and oil phase contamination of the water samples
shows a dramatic effect. In Figure 12, space velocity is presented
leads to great variation in the results. Samples representing part
in terms of volume (liquid hourly space velocity). By this
and all of the oil product for two-stage upflow tests are given.
measure, the new downflow results in this paper stand out
The effect of biocrude feed properties on the reactivity can
clearly compared to all of the earlier-published results. Thus,
be significant. Whereas the work at PNNL over the years had
our earlier agreement with Veba that the effect of downflow
used a number of bio-oils including several hardwood oils from
operation compared to upflow was small considering the overall
different fluid-bed pyrolysis reactors, pine oil from ablative
reaction and little difference between our earlier results and those
pyrolysis, hardwood oil from vacuum pyrolysis, and peat oil
from fluid-bed pyrolysis without identifying any major differ-
(39) Baldauf, W.; Balfanz, U. Upgrading of Pyrolysis oils from Biomass
ences in reactivity; more recent work showed large differences in Existing Refinery Structures; final report JOUB-0015, Veba Oel AG:
between poplar oil from ablative pyrolysis (NREL) and euca- Gelsenkirchen, 1992.
1804 Energy & Fuels, Vol. 21, No. 3, 2007 Elliott

Figure 9. Relation of oil viscosity to oxygen content.

Figure 10. Effect of temperature on oil viscosity.

Table 13. Low-Severity Hydrotreating Results in a Downflow Operation


TK751 catalyst K8-11
1st temp, °C 150 148 148 150 150 148 148 148 157
2nd temp, °C 380 375 354 349 349 362 355 400 ( 15 435 ( 45
LHSV, L/(L h) 0.28 0.30 0.29 0.29 0.29 0.43 0.38 0.38 0.31
WHSV, g/(g h) 0.52 0.54 0.53 0.53 0.54 0.78 0.70 0.64 0.51
yield, g/g biocrude 0.42 0.39 0.38 0.44 0.49 0.47 0.53 0.21 0.21
deoxygenation, % 98.6 97.9 96.3 95.1 95.0 94.5 95.8 97.6 97.1
density, g/mL 0.82 0.84 0.84 0.86 0.86 0.86 NA NA NA
gasification, % C 27 20 22 25 33 29 29 23 22
H2 consumption, L/L 881 746 727 808 813 791 779 494 313
carbon balance, % 92 79 79 91 107 101 109 54 53
Gas Effluent Composition (Including Excess Hydrogen)
hydrogen 70.3 77.4 78.4 70.7 72.2 73.3 73.4 77.8 85.8
carbon dioxide 8.1 7.0 5.8 7.6 5.5 5.5 7.1 7.5 5.8
methane 11.2 9.9 8.0 9.8 7.5 7.1 8.0 9.4 4.4
ethane 4.7 4.2 3.9 4.8 3.8 3.5 4.0 1.7 1.2
propane 2.1 0.4 1.2 2.2 3.6 3.5 2.1 1.0 0.8
butanes 1.4 0.3 0.9 1.8 2.6 2.4 1.5 0.6 0.5
pentanes 1.0 0.3 0.8 1.5 2.2 2.0 1.2 0.5 0.3
higher HCs 1.2 0.5 1.0 1.6 2.5 2.6 2.7 1.6 1.2

of Veba. However, Figure 13 provides a data comparison on a downflow results are much improved compared to our earlier
weight basis (weight hourly space velocity) wherein the effect upflow results. In all of these tests, the catalysts are supported
of the diluted CoMo catalyst bed used by Veba has a dramatic on an alumina base. The differences in the use of CoMo versus
effect. By using the diluted CoMo catalyst bed, Veba achieved NiMo are relatively small compared to the difference between
much higher processing rates based on the weight of catalyst. upflow and downflow.
The one unexplained result is the poor showing of the NiMo The product-oil density is clearly a function of product-oil
catalyst in the downflow experiments at Veba. Again, our oxygen content as seen in Figure 14. The molecular weight may
DeVelopments in Hydroprocessing Bio-oils Energy & Fuels, Vol. 21, No. 3, 2007 1805

Figure 11. Relation of carbon in the aqueous phase to the oxygen content of the oil.

Table 14. Low-Severity Hydrotreating Results with Different of bio-oil in a refinery draws the inappropriate conclusion that
Bio-oils product from 50 pyrolysis plants would be required to displace
NREL Union Fenosa only 5% of a conventional petroleum refinery. The conclusion
temperature, °C 355 365 is based on the faulty assumption that the pyrolysis plant would
WHSV, g oil/(g catalyst h) 0.70 0.54 only operate at only 40 t/d of biomass feedstock whereas
yield, g/g biocrude 0.53 0.41 biomass conversion plant concepts are typically at least an order
deoxygenation, % 96 92 of magnitude larger and a single 2000 t/d plant, which is the
product density, g/mL 0.86 0.94
gasification, % carbon 29 17 concept used by the Department of Energy in the U.S., would
hydrogen consumption, L/L oil 779 554 provide the 5% displacement.
carbon balance, % 109 85 After working with bio-oil,39 Veba’s conclusion was that, due
to its immiscible nature, it was impossible to process bio-oil
also be correlated with the oxygen content as they may both be together with petroleum products without pretreatment. Fol-
changing coincidentally as a function of processing severity. lowing pretreatment, there are a number of possible sites where
Figure 14 shows that the function of oxygen content and density the upgraded bio-oil could enter the refinery process streams.
seems to apply to the full range of products reported in the They performed catalytic hydrotreating as the upgrading process
literature as well as those from these tests. An important to make bio-oil (RTP from Ensyn, filtered to remove the 2%
conclusion to draw from these data is that the product-oil density char) compatible with petroleum. They used both sulfided CoMo
approaches 1 at a relatively low oxygen content, about 10%. and sulfided NiMo catalysts in a continuous-feed bench-scale
Products with oxygen contents around 10-15% and densities reactor operated at 17.8 MPa. At weight hourly space velocities
around 1 tend to form emulsions with the water byproduct and of 0.25-0.8 g/(g h) and temperatures of 350-370 °C, deoxy-
cannot be easily separated from the water. This lack of genation rates of 88-99.9% were achieved. Yields were
separation defeats one important purpose of the hydrotreating relatively constant at 30-35% oil, 50-55% water, and 15-
which is to remove the water and, thereby, dramatically improve 20% gases. Hydrogen consumption ranged from 350 to 600
the energy content of the oil. Nm3/t (420-720 L/L), 50-75% of which was used in water
Veba Oel AG. Veba Oel performed a series of handling tests formation. At the high severity of 99.9% deoxygenation, the
and upgrading experiments with biomass fast pyrolysis oil to oil consisted of 40% naphtha, 40% middle distillates, and 20%
evaluate its potential as a petroleum refinery feedstock Their vacuum-gas oil and was a yellow to white clear liquid. The
initial assessment of the use of bio-oil in a petroleum refinery heating value of the liquids was increased from 14.5 MJ/kg for
was evidently prepared before they had a useful understanding the bio-oil to 40-44 MJ/kg for the upgraded oil. At severities
of the material, as they suggest either putting the bio-oil into a below 95% deoxygenation, the product properties are unac-
desalting unit or distillation columns.40 Neither option is ceptable for introduction into the petroleum refinery.41
reasonable, as the water-wash step, which is desalting, would Veba reported that the catalyst deactivated quickly within 8 d.
dissolve more than half of the bio-oil into the water and the Gumlike deposits were found in the feeding tubes, valves, flow
bio-oil would not distill in the fractionating columns but would meters, and level controllers. At lower severity, the upgraded
polymerize to solids. In addition, the assessment for utilization
(41) Baldauf, W.; Balfanz, U. Upgrading of Fast Pyrolysis Liquids at
(40) Rupp, M. Utilisation of Pyrolysis Liquids in Refineries. Biomass Veba Oel AG. Biomass Gasification and PyrolysissState of the Art and
Pyrolysis Liquids Upgrading and Utilization; Bridgwater, A. V., Grassi, Future Prospects. Kaltschmitt, M., Bridgwater, A. V., Eds.; CPL Press:
G., Eds.; Elsevier Applied Science: London, 1991; pp 219-225. Newbury, UK, 1997; pp 392-398.
1806 Energy & Fuels, Vol. 21, No. 3, 2007 Elliott

Figure 12. Deoxygenation on a liquid hourly space velocity basis.

Figure 13. Deoxygenation on a weight hourly space velocity basis.

Figure 14. Relation of product oil density to oxygen content.

oil became more and more soluble in water, which led to sepa- tral thermowell), using both upflow and downflow configura-
ration problems. With the CoMo catalyst, the reactor blocked tions.42 The operating conditions tested and properties of some
in the preheating portion of the catalyst bed, and with NiMo, of the products are given in Table 15. Veba concluded that correl-
the coking was found in the outlet, not in the catalyst bed. A ating with a decreasing deoxygenation rate there was an increase
fixed-bed catalyst did not seem to be very practical and ebullated in specific gravity (see Figure 14), an increase in dissolved water
beds or liquid-phase systems (without catalyst) were suggested.
(42) Baldauf, W.; Balfanz, U.; Rupp, M. Upgrading of Flash Pyrolysis
Upgrading Tests. The upgrading tests were performed in a tub- Oil and Utilization in Refineries. Biomass Bioenergy 1994, 7 (1-6), 237-
ular reactor, 30 mm i.d. by 1.13 m long (0.716 L minus the cen- 244.
DeVelopments in Hydroprocessing Bio-oils Energy & Fuels, Vol. 21, No. 3, 2007 1807

Figure 15. Relation of dissolved water content in the product oil to the level of deoxygenation.

Table 15. Test Conditions and Properties of Upgraded Products


catalyst KF-702 C424
flow mode down down down down down down up down
temperature, °C 370 370 350 350 370 370 370 370
space velocity, kg/(kg h) 0.25 0.4 0.25 0.4 0.8 0.4 0.4 0.1
operating pressure, MPa 17.8 17.8 17.8 17.8 17.8 17.8 17.8 17.8
chemical H2 consumption 5.01 4.42 4.10 3.68 3.16 4.12 4.01 5.01
deoxygenation, % 99.92 98.53 95.17 91.78 88.32 95.88 95.25 97.20
Elemental Analysis, Water Free, wt % (Normalized to 100%)
carbon 86.79 87.22 86.84 86.5 85.94 86.93 86.82 86.47
hydrogen 13.17 12.35 11.82 11.23 10.79 11.87 11.52 12.90
oxygen 0.021 0.39 1.25 2.14 3.06 1.06 1.2 0.6
sulfur 0.07 0.004 0.004 0.008 0.055 0.005 0.007 0.021
nitrogen 0.003 0.033 0.082 0.121 0.152 0.129 0.23 0.009
density, g/mL 0.83 0.86 0.896 0.928 0.94 0.90 0.931 0.866
water content, ppm 50 210 510 1900 4200 350 351 102
CCR, wt % 0 0.09 1.40 3.39 5.52 1.79 1.60 0.28
aromatic carbon, wt % 10.0 17.0 18.8 25.9 27 18.8 22.7 9.8
HHV, MJ/kg 45.96 45.26 44.46 43.43 41.2 44.38 43.44 44.83
LHV, MJ/kg 43.09 42.6 41.87 40.99 38.87 41.68 40.91 42.00
Boiling Range, wt %
0-220 °C 40.8 40 30.6 27.6 25.3 29.8 21.4 38.5
200-350 °C 38.5 36.2 31.4 29.6 29.2 30.2 32.1 36.3
350-500 °C 20.7 21.2 23.1 23.0 21.6 23.8 27.6 22.3
>500 °C 2.6 14.9 19.8 23.9 16.2 18.9 3.0

(see Figure 15), a decrease in heating value, an increase in Conrad- Detailed analysis of the products was also undertaken. As
son carbon residue (CCR; see Figure 16), an increase in aromatic shown in Table 16, several of the product fractions as well as
content (see Figure 17), and a decrease in gasoline and middle the whole product oil were analyzed. From these analyses, it is
distillates while vacuum-gas oil remained fairly constant. concluded that the naphtha fraction has low octane because it
The influence of the different catalysts and flow mode seemed contains high levels of cyclic compounds which will require
to be as follows: reforming before it can be used in the gasoline pool at any more
• Under equal conditions, the deoxygenation in the upflow than small proportions; the diesel fraction is high density and
is lower than that in the downflow. low cetane because of a high aromatic content which will require
• Deoxygenation with the NiMo catalyst is less than with hydrogenation to stabilize it; the vacuum-gas oil (VGO) meets
the CoMo catalyst. the limits for CCR and metals content for use in hydrocracking;
• Depending on deoxygenation, lower amounts of naphtha and the residue fraction is a very small portion which might
are formed in upflow while the highest amount of vacuum gas not be separated in conventional distillation. It was recom-
oil is formed. mended that the total product be routed to the crude distillation
• The aromatic carbon content seemed to be higher in upflow. tower, where the fractions will end up diluted and upgraded in
• With the NiMo catalyst, there is a lower aromatic carbon subsequent conversion units.
content. Corrosion Tests. Corrosion problems were attributed to the
Evidence of the catalyst deactivation is seen by comparison high levels of organic acids in the bio-oil. Veba concluded that
of data sets 2 and 6 at the same conditions but 150 h later, i.e., only special types of steels and packings would be suitable for
the deoxygenation has dropped from 98.53 to 95.88. processing plants. Corrosion tests were performed at 94 °C with
1808 Energy & Fuels, Vol. 21, No. 3, 2007 Elliott

Figure 16. Relation of Conradson carbon residue in the product oil to the level of deoxygenation.

Figure 17. Relation of aromatic content in the product oil to the level of deoxygenation.

the type 1.4541 FeCrNi-steel, the lowest grade steel used in stabilization and direct combustion and/or extraction of chemi-
the plants. The weight losses of the steel samples observed after cals should be the focus of future activities. At that time (1994),
7 d of treatment amounted to 500 ppm with a corresponding the authors did not consider pyrolysis oil to be an economically
rate of 70 ppm/d. The concentration of Fe, Cr, and, Ni increased attractive feedstock for a petroleum refinery, because the
from 69, 2, and 4 ppm in the bio-oil to 380, 210, and 47 after upgraded bio-oil cost would be from 300 to 800 ECU/t while
7 d. At room temperature, no corrosion could be measured. petroleum and petroleum product costs were between 100 and
During the upgrading tests, corrosion was observed on the 200 ECU/t.
thermowell (type 1.4570 steel). After changing to type 1.4980, DMT FuelTec. A pilot-scale test of a slurry-phase hy-
an acid-resistant steel, there was less corrosion found. drotreating technology was performed in 1997 by DMT.43 The
Stabilization Tests. Batch reactor hydrogenations of the bio- Kohleoel pilot plant using the integrated gross oil refining
oil were performed at low temperature to evaluate the effect of (IGOR) technology was used for processing the heavy pyrolysis
a Pd catalyst for olefin removal. Through the use of NMR, a oil from the Union Fenosa plant. The bio-oil was processed
direct correlation of temperature from 25 to 155 °C with olefin with powdered NiMo catalyst at 30 MPa and 380 °C in a 2 m
removal was found; however, the overriding conclusion was
that this type of processing led to a heavy bitumenlike product. (43) Kaiser, M. Upgrading of Fast Pyrolysis Liquids by DMT. Biomass
Gasification and PyrolysissState of the Art and Future Prospects;
Some water separation was also accomplished, but handling was Kaltschmitt, M., Bridgwater, A. V., Eds.; CPL Press: Newbury, UK, 1997;
difficult. However, Veba concluded that cheap alternatives for pp 399-406.
DeVelopments in Hydroprocessing Bio-oils Energy & Fuels, Vol. 21, No. 3, 2007 1809

Table 16. Characteristics of Hydrotreated Bio-oil Products


naphtha diesel VGO residue
total product <180 °C 180-350 °C 350-500 °C >500 °C
Elemental Analysis
carbon, wt % 86.45 85.45 87.08 87.54
hydrogen, wt % 13.22 14.72 12.88 12.26
oxygen, wt % 0.2 0.05 0.07 0.4
sulfur, ppm 851 130 109 426
nitrogen, ppm 83 2 51 162
calcium, ppm <1
nickel, ppm <1
vanadium, ppm <1
molybdenum, ppm <1
iron, ppm <1
sodium, ppm <1
density, g/mL 0.8503 0.7517 0.8739 0.9303 0.9316
n-paraffins wt % 20.56
iso-paraffins, wt % 12.52
naphthenes, wt % 61.41
polynaphthenes, wt % 1.47
aromatics, wt % 4.03
monoaromatics, wt % 29.3
diaromatics, wt % 0.6
polycyclic aromatics, wt % 0.7
MON/RON 62/62
cetane number 45
cetane index 39.7
cloudpoint, C -34
CCR, wt % 0.02
stability test, ASTM D-2274 negative

Table 17. Properties of Bio-oil and Hydrotreated Products and Hydroprocessing Results
bio-oil feed Co-Mo 3.0%-14.3% Ni-Mo 2.9%-15.1% NiMo w/tetralin
carbon, wt % 73.7 84.0 84.4 86.9
hydrogen, wt % 7.6 8.0 9.5 10.4
oxygen, wt % 15.3 5.5 4.7 2.2
nitrogen, wt % 3.3 2.5 1.4 0.5
sulfur, wt % 0.1 not reported not reported not reported
H/C atomic ratio 1.23 1.13 1.35 1.44
moisture content, wt % 9 not reported not reported not reported
viscosity, cPs @60C 480 not reported not reported not reported
density, g/mL 1.19 not reported not reported not reported
pH 5.5 not reported not reported not reported
solubility in THF 85 not reported not reported not reported
BP < 370 °C not reported 50% 95%
gas yield, wt % 10 12-15
oxygen removal, % 64 70 85
nitrogen removal, % 24 58 85
hydrogen consumption, L STP/L oil 380 440 a
a 85% of the tetralin was recovered after the test.

by 10 cm reactor tube. The process includes subsequent Raiano plant. In particular, they focused on a reduction of the
separations and fixed-bed catalytic process steps resulting in level of nitrogen (3.3 wt % in the bio-oil) in order to reduce
three product streams of syncrude, refined syncrude, and highly NOx emissions in the combustion of the bio-oil. They used
refined syncrude. A yield of 33.9% of the lightest product was sulfided CoMo and NiMo catalyst formulations based on
claimed, and the analysis given showed a highly deoxygenated alumina support. Dimethyldisulfide was added to maintain an
product containing less than 0.1% oxygen, no coke, ash or water, H2S partial pressure of 0.1-0.2% at reaction conditions. In
a density of 0.801, and a distillation range from 40 °C IBP, batch-reactor tests, temperatures from 300-400 °C at 12 MPa
50% at 263 °C and 95% at 349 °C. The other two liquid products of reactor pressure (hot) were used with residence times of 1 to
accounted for 9.1% yield, but no analysis was provided. 2 h. These tests showed promise for heteroatom removal by
Hydrogen consumption was given at 9.1 kg/kg with 1.0 kg this method. Results with the two catalysts are given in Table
product gas which could potentially be reused. Another 25.2% 17.
yield of hydrocarbon gases was produced. The throughput in The authors reported that only sulfided NiMo or CoMo on
the slurry reactor was at 1 LHSV; however, the sizes of the alumina or silica-alumina supports would work for this
subsequent fixed-bed reactors were not provided, so an overall reaction. They claim that supported noble metal catalysts would
space velocity in the catalytic system cannot be determined. be readily deactivated through chemical and physicochemical
No attempt was reported to determine costs for such processing. processes such as poisoning, sintering, and fouling. The authors
Université Catholique de Louvain. During the 1990s, the
Catholic University of Louvain in Belgium carried on a (44) Churin, E.; Maggi, R.; Grange, P.; Delmon, B. Characterization and
significant laboratory research effort in catalytic hydrogenation Upgrading of a Bio-Oil Produced by Pyrolysis of Biomass. Research in
Thermochemical Biomass ConVersion; Bridgwater, A. V., Kuester, J. L.,
for improving the quality of bio-oil. The initial effort44 focused Eds., Elsevier Science Publishers, LTD. Barking, England.: 1988; pp 896-
on a bio-oil produced by the pyrolysis of olive oil waste at the 909.
1810 Energy & Fuels, Vol. 21, No. 3, 2007 Elliott

relate that the lowest viscosity and oxygen-content product was


produced at 300 °C as opposed to higher-temperature tests at
350 and at 400 °C. However, they modified their method to
include a low-temperature stabilization at temperatures below
300 °C followed by processing at 360-380 °C, as reported by
Elliott and Baker.29 Processing was also performed with a
hydrogen-donor solvent, tetrahydronaphthalene, and improved
processing was demonstrated, as seen in Table 17 wherein
tetralin was added in a 1-to-1 ratio with bio-oil feedstock.
There are two special conditions in these testss(1) the bio-
oil contained a large amount of nitrogen, and significant levels
of pyridines and indoles were found in the products; and (2)
the bio-oil contained a large fraction of hydrocarbons (from the
olive oil), and therefore, large amounts of hydrocarbon products
were found. Lignocellulosic biomass typically contains neither
of these two groups.
In a report of the broader scope of this study,45 it was
concluded that

if the pyrolysis oil is hydrotreated alone, the extent of Figure 18. Reactivity scale of oxygenated groups under hydrotreatment
the improvement is limited because of the polymeriza- conditions.
tion of the compounds able to generate free radicals
which are not saturated rapidly enough and because the the aim is to focus specifically on using hydrogen for oxygen
high oxygen concentration leads, after hydrodeoxygen- removal, then the much higher selectivity of the CoMo catalyst
ation, to high concentrations of water which deactivates for hydrodeoxygenation could be important. These model
the catalysts by transforming the active sites. compound tests also showed that these catalysts had significant
sChurin et al. activity at mild conditions of 300 °C and 30 bar.
A more detailed study was undertaken of the catalytic hydro-
The use of hydrogen donor solvent leads to a marked genation reactions using model compounds to evaluate sulfided
improvement of the quality of the product, and the catalysts NiMo and CoMo catalysts and the inhibitions of water, hydrogen
are less deactivated by coke deposition. The authors reported sulfide, and ammonia, the products of the hydrotreating reac-
that distillation of the product could be used to recover a fraction tions.48 The work was done in a stirred batch rector using a dilut-
which could serve as the hydrogen-donor solvent after it was ing solvent (dodecane) to minimize the polymerization reactions
hydrogenated. so that conclusions could be made about the relative importance
Subsequent tests with a bio-oil produced from acacia wood of the two hydrogenation reaction pathways: hydrodeoxygen-
in the S.A. Bio-Alternative plant in Switzerland were used to ation and ring saturation (hydrogenation) of phenolic compounds.
define the means to control the extent of hydrotreating.46 Batch- Both catalysts were commercial products (Procatalyse HR 306
reactor tests using a sulfided NiMo catalyst (Procatalyse HR346) and HR 346) based on γ-alumina supports. Tests with para-
with donor solvent diluted bio-oil were performed at 200 and cresol showed that, with the more active hydrogenation catalyst,
350 °C for 140 min. On the basis of the results and other NiMo, as well as the more active hydrodeoxygenation catalyst,
information from the literature about the reactivity of oxygenated CoMo, the inhibitions are similar.49 The inhibition strength
groups over sulfided catalysts,47 a reactivity scale was developed follows the order water < 2-ethylphenol < hydrogen sulfide <
as shown in Figure 18. As shown, at low temperatures, olefins, ammonia. At low concentrations, hydrogen sulfide promotes
aldehydes, and ketones are readily reduced by hydrogen. These slightly the hydrogenation activity of the CoMo catalyst which
reactions stabilize the bio-oil by removing these well-known is contrary to the inhibiting effect of the compounds and its
reactive groups. Alcohols are reacted at 250-300 °C by catalytic effect on NiMo. Since sulfide is present only at low levels in
hydrogenation but also by thermal dehydration to form olefins. bio-oils, it will likely need to be added to maintain the sulfided
This olefin formation may lead to the difficulties with bio-oil character of these catalysts but its inhibiting effect at higher
polymerization at hydrotreatment conditions. Carboxylic and concentration indicates that its addition will need to be carefully
phenolic ethers are also reacted at around 300 °C. Di-phenyl controlled. Later studies50 with carbon-supported CoMo catalyst
ether and dibenzofuran have not been found as significant confirmed that the hydrogen sulfide caused inhibition (above a
components in bio-oil. 75 kPa partial pressure of H2S) of dehydroxylation of guaiacol
Model Compound Studies. On the basis of model compound favoring a selectivity shift toward catechol production and also
tests35 involving o-cresol and naphthalene, it was concluded that an inhibition of carbonyl hydrogenation. The inhibition was less
NiMo with a phosphated alumina support was the most active marked in the conversion of carboxyl groups.
for oxygen removal and hydrogen incorporation. However, if
(48) Laurent, E. Etude et contrôle des réactions d’hydrodésoxygénation
lorts de l’hydroraffinage des huiles de pyrolyse de la biomasse. Doctor of
(45) Churin, E.; Grange, P.; Delmon, B. Quality ImproVement of Pyrolysis Sciences dissertation, Université Catholique de Louvain, Louvain-la-Neuve,
Oils; final report on contract no. EN3B-0097-B for the Directorate-General Belgium, 1993.
Science, Research and Development, Commission of the European Com- (49) Laurent, E.; Delmon, B. Influence of Oxygen-, Nitrogen-, and Sulfur-
munities, 1989. Containing Compounds on the Hydrodeoxygenation of Phenols over
(46) Laurent, E.; Pierret, C.; Grange, P.; Delmon, B. Control of the Sulfided CoMo/R-Al2O3 and NiMo/R-Al2O3 Catalysts. Ind. Eng. Chem. Res.
Deoxygenation of Pyrolytic Oils by Hydrotreatment. Presented at Biomass 1993, 32, 2516-2524.
for Energy, Industry and Environment, the 6th EC Conference, Athens, (50) Ferraari, M.; Bosmans, S.; Maggi, R.; Delmon, B.; Grange, P. CoMo/
Greece, April 22-26, 1991. Carbon Hydrodeoxygenation Catalysts: Influence of the Hydrogen Sulfide
(47) Weisser, O.; Landa, S. Sulfided Catalysts, Their Properties and Partial Pressure and of the Sulfidation Temperature. Catal. Today 2001,
Applications. Pergamon Press: New York, 1973. 65, 257-264.
DeVelopments in Hydroprocessing Bio-oils Energy & Fuels, Vol. 21, No. 3, 2007 1811

In a second series of tests, mechanistic pathways were Table 18. Low-Temperature Catalytic Hydrotreating of Ensyn
evaluated for oxygenated compounds of interest.51 A solution Bio-Oil
of guaiacol, methyl-acetophenone, and diethyl decanedioate in bio-oil feedstock upgraded bio-oil
hexadecane solvent was hydrogenated over a temperature range Elemental Analysis
from 260 to 300 °C at 7 MPa with hydrogen overpressure. It C, wt % 52.9 not reported
was found that the ketone group in methylacetophenone was H, wt % 6.2 not reported
easily hydrogenated with either catalyst at temperatures higher N, wt % 0.2 not reported
O, wt % (by difference) 40.7 not reported
than 200 °C. Ethylmethyl benzene was quantitatively produced acetone insoluble, wt % 2.98 not determined
with no intermediates found. The NiMo catalyst has a better ash, wt % 0.20 0
activity for the conversion of guaiacol than the CoMo catalyst. neutralization number, 100 5.0
Catechol is the primary product with phenol being the secondary mg KOH/g
pH 3.17 6.5
product. Subsequent products of benzene and cyclohexane were moisture content, wt % 24.8 34.3
also seen at long reaction times. The CoMo catalyst has a density, g/mL@20 °C 1.21 0.98
slightly better selectivity for catechol + phenol while the NiMo viscosity, cPs @ 20 °C 179 not determined
led to more side reactions with unidentified products. The diester HHV (MJ/kg, m.f.) 19.4 not reported
n-pentane solubles, wt % 0.25 47.89
went through a series of reactions with products including
asphaltenes (benzene 4.27 35.60
monoesters, diacids and monoacids, and finally the alkane soluble), wt %
hydrocarbons. The formation of C9 products suggested some residue, wt % 95.48 16.51
decarboxylation reactions as well as the hydrogenolysis reac-
tions. The NiMo had a slightly higher activity. mm i.d. × 624 mm length) constructed of 304 SS was operated
The same solution of model compounds was used to evaluate in an upflow configuration. Initial testing was done with a model
the inhibition of catalysis by ammonia, water, and hydrogen compound solution of acetophenone, guaiacol, dibenzofuran,
sulfide, the products of hydrotreating of heteroatoms.52 Similar di-ethylsebacate, and 2-ethylphenol, each at 10 wt % in an
to the results described above with phenolic compounds, water aliphatic solvent carrier. The catalyst was the Akzo KF-840
had little effect on the reactions; ammonia had quite a strong NiMo catalyst, and the feed solution included dimethyldisulfide
effect except for the hydrodeoxygenation of ketones; and for a 0.2 wt % solution to maintain the sulfided form of the
hydrogen sulfide depresses the activity of NiMo for conversion presulfided catalyst. The test was performed at 280 °C and 10
of ketones but not for CoMo, and the demethylation of guaiacols MPA with a liquid hourly space velocity of 1 and a hydrogen-
was not affected. The authors claimed that these results to-feed ratio of 1000 vol/vol. In this test, the model compounds
suggested that hydrogen sulfide and ammonia may be used for were 100% converted except for the dibenzofuran, which was
controlling the reactions occurring in hydrotreating of bio-oils. 94% converted, but products were not reported. Subsequent tests
under similar conditions with bio-oil actual feedstocks (Ensyn
A more detailed study53 of the effect of water in the hydro-
and Union Fenosa) were less successful with plugging in the
treating system was undertaken to evaluate longer-term effects.
inlet and in the front part of the catalyst bed after a few hours
Using model compounds, para-cresol and dibenzofuran, pro-
of operation.56 A table of data from one test was reported (see
cessed in dodecane solvent, continuous-flow reactor tests were
Table 18). The operating temperature was reported at 280 °C
performed at 340 and 360 °C at 7 MPa. In this study, it was
with an exotherm which caused a bed profile from 280 to 306
calculated that for typical hydrotreating conditions of bio-oil
°C before falling back to 280 °C at the outlet. The LHSV was
the water vapor pressure is 2-3 orders of magnitude higher than
reduced to 0.5 while maintaining the same hydrogen flow such
that found in a typical petroleum hydrotreater. The treatment
that the relative flow increased to 2000 vol/vol at an operating
studies found that a NiMo catalyst in the presence of this amount
pressure of 14 MPa. The oil product shows evidence of
of water vapor loses two-thirds of its activity in less than 60 h
hydrotreating, but the reported analyses are surprising in a few
of treatment independent of the hydrogen sulfide partial pressure.
respects: the neutralization number is low as is the density,
It was concluded that a permanent structure change occurred as
and they do not fit with the high moisture content which suggests
the catalyst activity could not be recovered. The major change dis-
a large residual oxygen content.
covered was the conversion of R-alumina to the hydroxide form,
Modifications to the reactor system, including a cooler on
boehmite. As a result, there was a small decrease in surface
the inlet side of the reactor, allowed longer-term operations.57
area and a concurrent sintering of the nickel (but not the molyb-
denum). In addition, an oxidation of the nickel was detected and (52) Laurent, E.; Delmon, B. Study of the Hydrodeoxygenation of
the result was interpreted as either the formation of a noncata- carbonyl, carboxylic and gauaicyl groups over sulfided CoMo/γ-Al2O3 and
lytic nickel sulfate which could form a layer on the nickel and NiMo/γ-Al2O3 catalysts. II. Influence of water, ammonia and hydrogen
molybdenum sulfides further inhibiting the catalytic activity or sulfide. Appl. Catal. A 1994, 109, 97-115.
(53) Laurent, E.; Delmon, B. Influence of Water in the Deactivation of
the formation of noncatalytic nickel aluminates. The addition a Sulfided NiMo/γ-Al2O3 Catalyst during Hydrodeoxygenation. J. Catal.
of the model compounds to the reaction mixture did not modify 1994, 146, 281-291.
the results, suggesting that water is the strongest oxidizing agent. (54) Conti, L.; Scano, G.; Boufala, J. Bench Scale Plant for Continuous
Hydrotreating of Oils from Biomass. AdVances in Thermochemical Biomass
University of Sassari. A bench-scale reactor system for ConVersion Bridgwater, A. V., Ed.; Blackie Academic & Professional:
catalytic hydrotreatment of bio-oil was assembled and operated London, 1993; Vol. 2, pp 1460-1464.
at the University of Sassari during the 1990s. The system was (55) Conti, L.; Scano, G.; Boufala, J.; Mascia, S. Bio-Crude Hydrotreating
in a Continuous Bench-Scale Plant. DeVelopments in Thermochemical
designed to process up to 2 L/h at up to 450 °C and 20 MPa,54 Biomass ConVersion; Bridgwater, A. V., Boocock, D. G. B., Eds.; Blackie
although it was only operated at 140-280 °C and 15 MPa to Academic & Professional: London, 1996; Vol. 1, pp 622-632.
study the first step of bio-oil stabilization.55 The reactor (44 (56) Conti, L.; Scano, G.; Boufala, J. Continuous Hydrotreating Research
at University of Sassari. Proceedings of the Contractors Meeting AIR CT92-
0216, Gelsenkirchen, Germany, August 30-13, 1994.
(51) Laurent, E.; Delmon, B. Study of the Hydrodeoxygenation of (57) Conti, L.; Scano, G.; Mascia, S.; Caria, V. Improvement of the Flash-
carbonyl, carboxylic and gauaicyl groups over sulfided CoMo/R-Al2O3 and Pyrolysis Process and Pilot Plant for Bio-Oils Upgrading. Sixth Interim
NiMo/R-Al2O3 catalysts. I Catalytic Reaction Schemes. Appl. Catal. A 1994, Report; contract no. AIR-CT-92-0216; Energy Research Group, Aston
109, 77-96. University: Birmingham, U.K.; November 1996.
1812 Energy & Fuels, Vol. 21, No. 3, 2007 Elliott

Table 19. Low-Temperature Catalytic Hydrotreating of Union Fenosa Bio-Oil


bio-oil feedstock upgraded bio-oil
hours on stream 7.5 12
LHSV 0.51 0.55
hydrogen feed vol/vol 2228 2048
bed temperatures
bed entrance 114 116
bed middle 254 255
bed exit 273 275
liquid products, g/h 45.7 58.2
H2 consumption, L/kg 207
liquid recovery, wt % 80.2 93.9
deoxygenation, % 45
Elemental Analysis
C, wt % 51.6 62.2
H, wt % 6.2 7.5
N, wt % 0.5 0.5
O, wt % (by difference) 41.7 29.8
acetone insoluble, wt % 2.62 0.31
ash, wt % 0.11 0.03
neutralization number, mg KOH/g 85.6 5.0
pH 3.43 6.5
moisture content, wt % 16.7 3.0
density, g/mL@20 °C 1.27 1.07
viscosity, cPs @ 20 °C 1150 not reported
HHV (MJ/kg, m.f.) 19.3 26.4
n-pentane solubles, wt % not reported 21.1 (65.2 C, 7.7 H, 0.3 N, 26.8 O)
benzene solubles, wt % not reported 30.4 (66.7 C, 7.8 H, 0.4 N, 29.8 O)
residue, wt % not reported 48.5 (58.2 C, 6.3 H, 0.5 N, 34.0 O)

Table 20. Catalytic Hydrotreating of Ensyn Bio-Oil with a A final extended test was reported which operated for 120 h
Carbon-Supported Catalyst and was shut down voluntarily with no apparent deactivation
bio-oil feedstock upgraded bio-oil of the sulfided KF-840 catalyst. A bed temperature profile was
hours on stream 21 34.7 maintained with an entrance temperature from 125 to 148 °C,
LHSV 0.59 0.90 a midpoint temperature of 243-268 °C, and an exit temperature
hydrogen feed vol/vol 2228 1461 of 262-291 °C. The LHSV target was 0.5, and it varied from
bed temperatures 0.48 to 0.58. The hydrogen feed rate was constant at 105 L/h
bed entrance 150 140
bed middle 256 263 (2448-2013 vol/vol) at an operating pressure of 15 MPa. The
bed exit 247 248 overall product balance showed a liquid product yield of 72 wt
liquid products, g/h 50.6 79.2 % and a deoxygenation of 60 wt % requiring a hydrogen
H2 consumption, L/kg 132 consumption of 264 L/kg of bio-oil feed. The product oil
liquid recovery, wt % 93.2 95.7
deoxygenation, % 43.8
composition showed the higher level of deoxygenation with only
23.2 wt % oxygen. The pH and neutralization numbers are given
Elemental Analysis
C, wt % 54.1 65.0 64.4 as being identical to those from Tables 18 and 19 casting further
H, wt % 6.8 7.5 7.1 doubt on these measurements.
N, wt % 0.2 0.2 0.2 Université Catholique de Louvain (UCL)sCarbon Sup-
O, wt % (by difference) 38.9 27.3 28.3 ports. Extensive studies have been undertaken with carbon-
Results were reported for a test with Union Fenosa bio-oil using supported CoMo and NiMo catalysts as alternatives to conven-
the KF-840 catalyst in its sulfided form (dimethyldisulfide was tional alumina-supported petroleum industry catalysts. Initial
added with the bio-oil to generate a 0.2 wt % sulfur feedstock). studies showed that coke formation was an important element
The test was kept online for a longer period apparently because in catalyst deactivation with the use of alumina supports
of the lower temperature in the front of the reactor. The results principally with molecules containing two oxygens such as
are presented in Table 19. The operating pressure for the test guaiacols or catechols.58 Carbon-supported CoMo and NiMo
was 15.2 MPa. Curiously, the numbers for the pH and catalysts were formulated at UCL and tested with model
neutralization number are identical to the test results in Table compounds in a batch-reactor system.59 The model compounds
18. The neutralization number is surprisingly low given the high included guaiacol, catechol, phenol, 4-methyl acetophenone,
oxygen content remaining in the product oil. The higher density ethyl decanoate, and para-cresol tested in para-xylene solutions.
seems more reasonable. It was found that in general the catalytic activity for the
An extended test was also performed with a sulfided carbon- hydrodeoxygenation of carbonyl, carboxyl, and methoxyl groups
supported NiMo catalyst using the Ensyn bio-oil as the (3 times higher) was higher for alumina-supported catalysts
feedstock. After about 65 h, the experiment was stopped due compared to carbon-supported catalysts (both CoMo and NiMo).
to a high pressure drop across the reactor. The results from two An important difference is in the selectivity to phenol versus
of the data windows are presented in Table 20. The operating catechol with the use of carbon supports and the apparent
pressure for the test was 15.2 MPa. The gas composition was avoidance of coke formation as indicated by near stoichiometric
analyzed and the products were determined to be mostly carbon (58) Centeno, A.; Laurent, E.; Delmon, B. J. Catal. 1995, 154, 228.
dioxide and methane with some higher hydrocarbons. Carbon (59) Centeno, A.; David, O.; VanBellinghen, Ch.; Maggi, R.; Delmon,
monoxide was also detected with this catalyst whereas it was B. Behavior of Catalysts Supported on Carbon in Hydrodeoxygenation
Reactions. DeVelopments in Thermochemical Biomass ConVersion Bridg-
undetectable in the test with the alumina-supported catalyst (KF- water, A. V., Boocock, D. G. B., Eds.; Blackie Academic & Professional:
840). London, 1996; Vol. 1, pp 589-601.
DeVelopments in Hydroprocessing Bio-oils Energy & Fuels, Vol. 21, No. 3, 2007 1813

recovery of products with carbon supports as opposed to only work, the two-step concept was demonstrated using Ru catalyst
75% for alumina-supported CoMo. Further studies of the carbon- at very low temperature followed by a moderate hydrotreating
supported catalysts evaluated method modifications such as the by NiW at relatively low temperature to produce a product with
order of metal impregnation and intermediary thermal or less than 10% oxygen content.
sulfidation treatment.60 The conclusions of the study were that University of Hamburg. Some batch reactor results with
the activated carbon textural characteristics were not modified palladium, copper chromite, and nickel catalysts have been
by the intermediary treatments; both intermediary treatments reported.63 In those tests, operation at 20 °C caused slight
increased the stability of the metal sulfides but did not influence changes, while tests at 100 °C resulted in what were described
the dispersion; the MoCo catalysts (cobalt impregnated first, as “drastic changes”. Various ketones were removed, but acetic
followed by molybdenum) have better active phase dispersion, acid was not reduced. Although gas chromatographic separations
which favors hydrodeoxygenation activity for methoxy and were performed on the products, detailed chemical conversion
carboxy groups. There was an activity decrease for carbonyl conclusions were not reported.
groups, but the activity was still very high. There was a strong Hydrogenation of fast pyrolysis oil was also studied by
increase for hydrogenation selectivity versus decarboxylation. Scholze64 using a batch reactor with various metal catalysts and
A later paper61 by the same group concludes that the MoCo without catalysts at low temperatures. She concluded that
catalyst does not have a better dispersion but the metals are reaction temperatures above 80 °C are unsuitable for hydroge-
preferentially found on the surface of the catalyst. The study nation of bio-oils because the product phases separate. Further,
found that cobalt is preferentially deposited on the surface of none of the combinations of bio-oils catalysts and conditions
the catalyst while molybdenum penetrated more deeply into the tested resulted in a more stable oil. She found that palladium
pore structure. When the second metal was added, it tended to was essentially inactive at 60 °C. Raney nickel at 80 °C resulted
join the existing metal site, although there was some evidence in reduced viscosity over time (without phase separation), while
of molybdenum migrating out of the pores upon impregnation copper chromite at the same temperature resulted in a slightly
with the cobalt. In the case of the MoCo catalyst, the Mo is more viscous oil over time. Nickel metal was tested at
mainly impregnated at the exterior of the carbon particles, temperatures from 22 to 100 °C. At 22 °C, there was noticeable
forming a thick layer of oxide agglomerates, which probably reduction in carbonyl (∼15%) without a noticeable change in
trap and hide the Co. The authors conclude that cooperation physical properties. At 82 and 100 °C, the product oil separated
between the two components would be extremely difficult and into two phases (as did the copper chromite catalyzed product).
the promoting role of the Co would be almost nil. However, Chemical analysis of these products was performed to a limited
the carbon pore structure also interplays with these dispersions degree, but little was concluded about the changes in the oil
such that using a macroporous carbon allows the CoMo catalyst composition. Carbonyl analysis showed no change in the
to be more active because its active phase can be more highly palladium catalyzed tests and up to a 20% reduction at the
dispersed yet accessible, while the use of the microporous carbon intermediate temperature of 50 °C with nickel metal catalyst.
causes the MoCo formulation to be relatively more active, Although gas chromatographic separations were performed on
because its active phase is more accessible when concentrated the products, detailed chemical conversion conclusions were not
on the surface. reported.
Université Laval. Low-temperature hydrotreating was un- Pacific Northwest National LaboratorysPrecious Metal
dertaken using a 5% Ru on γ-alumina catalyst at temperatures Catalysis. Process development is underway on catalytic
as low as 80 °C.62 Copper chromite was also tested and found hydrotreatment for converting biomass fast pyrolysis oils to
to be not useful in this process as the bio-oil coked; the same upgraded liquid fuels and chemicals.65 Initial tests were
thing happened with no catalyst present. In the stirred batch performed using the white wood bio-oil and bagasse bio-oil from
reactor, vacuum pyrolysis bio-oil was slurried with the catalyst DynaMotive processed over a ruthenium on carbon catalyst.
and heated to temperature with a hydrogen overpressure. The process data are summarized in Table 21. Although several
Following a first period of 2 h, the reactor is cooled and opened, processing conditions were tested within the series, the attempt
a second hydrotreating catalyst (Ni/W/γ-Al2O3) is added, and was made to return to starting conditions to verify catalyst
the reactor is heated to 325 °C for another 2 h. The gas stability. Significant loss of catalyst reactivity was found.
consumption clearly shows that the bio-oil can be hydrogenated The product oils formed at these processing conditions are
even at these low temperatures with the Ru catalyst. The authors two-phased: an aqueous phase with higher hydrophilic com-
claim that the most effective combined hydrogenation-hy- ponents above a more dense tar phase composed of the more
drotreating is accomplished with 80 °C hydrogenation, based hydrophobic components with a lesser amount of dissolved
on elemental analysis of the products. In addition, the gel water. The tar-phase product oil analyses are shown in Table
permeation chromatography allowed molecular weight measure-
ments to be obtained which showed that polymerization occurred
(63) (a) Meier, D.; Wehlte, S.; Wulzinger, P.; Faix, O. Upgrading of
at temperatures as low as 140 °C even with hydrogen and a Bio-oils and flash pyrolysis of CCB-treated wood waste. Bio-Oil Production
catalyst present. Further, the hydrogenated product produced and Utilization; Bridgwater, A. V., Hogan, E. N., Eds.; CPL Scientific
at a lower temperature had a lower molecular weight and was Ltd: Newbury, UK, 1996; pp 102-112. (b) Meier, D.; Bridgwater, A. V.;
DiBlasi, C.; Prins, W. Integrated chemicals and fuels recovery from pyrolysis
more readily hydrotreated to lower oxygen content. By this liquids generated by ablative pyrolysis. Biomass Gasification and Pyroly-
sis: State of the Art and Future Prospects, Kaltschmitt, M., Bridgwater,
(60) Centeno, A.; VanBellinghen, Ch.; David, O.; Maggi, R.; Delmon, A. V., Eds.; CPL Scientific Ltd: Newbury, UK, 1997; pp 516-527.
B. Influence of the Preparation Procedure on Catalysts Supported on Carbon (64) Scholze, B. Long-term stability, catalytic upgrading, and application
in Hydrodeoxygenation Reactions. DeVelopments in Thermochemical Bio- of pyrolysis oilssImproving the properties of a potential substitute for fossil
mass ConVersion; Bridgwater, A. V., Boocock, D. G. B., Eds.; Blackie fuels. Doctoral dissertation, University of Hamburg, Hamburg, Germany,
Academic & Professional: London, 1996; Vol. 1, pp 602-610. 2002.
(61) Ferrari, M.; Delmon, B.; Grange, P. Influence of the Impregnation (65) Elliott, D. C.; Neuenschwander, G. G.; Hart, T. R.; Hu, J.; Solana,
Order of the Molybdenum and Cobalt in Carbon-supported Catalysts for A. E.; Cao, C. Hydrogenation of Bio-Oil for Chemical and Fuel Production.
Hydrodeoxygenation Reactions. Carbon 2002, 40, 497-511. Science in Thermal and Chemical Biomass ConVersion; Bridgwater, A. V.,
(62) Gagnon, J.; Kaliaguine, S. Catalytic Hydrotreatment of Vacuum Boocock, D. G. B., Eds.; CPL Press: Newbury Berks, UK, 2006; pp 1536-
Pyrolysis Oils from Wood. Ind. Eng. Chem. Res. 1988, 27, 1783-1788. 1546.
1814 Energy & Fuels, Vol. 21, No. 3, 2007 Elliott

Table 21. Bio-Oil Hydrotreating Process Data Table 23. Catalytic Hydrotreating of Pyrolytic Lignin
white wood bio-oil bagasse bio-oil temperature, °C, inlet/outlet 240/400 230/415
duration, h 3.03 1.57
LHSV, L bio-oil/(L catalyst h) 0.22-0.67 0.22-0.67
amount lignin fed, g 102.1 102.6
temperature, °C 181-238 182-240
space velocity, g of lignin/(g of catalyst h) 0.51 1.0
pressure, atm H2 133-142 132-143
deoxygenation, % 31-70 32-46 Yields, wt % of Dry Feed
oxygen balance 83-103 88-102 water 21.0 20.4
hydrogen consumption, L/L 261-491 163-311 light organics 64.1 60.5
product oil yield, dry g/g 0.54-0.79 0.64-0.81 total liquid product 85.1 80.9
heavy residue 1.6 1.5
Table 22. Bio-Oil Hydrotreating Product Analyses liquid + tar produced 86.7 82.4
white hydrogenated Gases
wood white wood bagasse hydrogenated carbon monoxide traces traces
bio-oil bio-oil bio-oil bagasse bio-oil carbon dioxide 3.30 3.66
methane 3.14 2.66
Product Oil Composition, wt % (Dry Basis) ethylene traces traces
carbon 48.3 65.6-74.0 52.0 65.1-70.7 ethane 0.96 0.83
hydrogen 7.4 8.5-10.1 6.6 7.7-9.4 propane 0.86 0.75
oxygen 44.4 16.7-25.8 41.3 20.3-27.0 butane and larger 0.91 0.75
H/C atomic 1.82 1.49-1.68 1.51 1.37-1.59 total gases 9.17 8.65
ratio (wet) total recovery, % 95.9 91.1
water content, 30.0 9.7-15.7 35.0 11.2-14.0
wt % Table 24. Properties of Bio-Oil Feedstock and Hydrotreated
Product
22. The aqueous-phase products contain from 18 to 27 wt % bio-oil feedstock hydrotreated product
carbon by total organic carbon analysis and are 48-61 wt %
density, g/mL 1.12 0.93
water by Karl Fischer assay. The aqueous-to-tar-product mass
ratio was about 6 to 1. Elemental Analysis
carbon, wt % 60.4 87.7
These product oils are complex mixtures like the feedstock hydrogen, wt % 6.9 8.9
bio-oils. The two phases appear to form not from polymerization oxygen, wt % 41.8 3
of the tar, as concluded by Scholze,64 but by the change in the nitrogen wt % 0.9 0.4
water solubilities of the product chemicals. As the products HHV, MJ/kg 21.3 41.4
solubility in methanol, wt % 99
become less hydrophilic by removal of the carbonyl, olefinic, solubility in toluene, wt % little 100
and aromatic characteristics, the delicate balance that allows a
single-phase product to form is upset and the two phases mixed with tetrahydronaphthalene, was used as the feedstock
separate. From these processing tests with the two feedstocks for catalytic hydroprocessing tests to evaluate temperature, resi-
at comparable conditions, the softwood-derived oil appears to dence time, and pressure.67 The catalyst was a CoMo (4.0/24
be more readily hydrogenated than the bagasse-derived bio- wt %) on phosphated (2.8%) γ-alumina. Properties of the feed-
oil. By extrapolation, the hardwood oil also appears to be more stock and a representative hydrotreated product are given in
difficult to hydrogenate because of its similarity in chemical Table 24. Deoxygenation as a function of temperature showed
structures to the bagasse-derived bio-oil. a consistent increasing effect at temperatures from 360 to 390 °C.
University of Waterloo. A pyrolytic lignin, extracted from Residence times in the batch reactor greater than 20 min had
softwood fast pyrolysis bio-oil, was upgraded by catalytic hydro- only minimal effect. The effect of pressure was little both
treating.66 The pyrolytic lignin was a heavy oil phase separated because the experiments were performed with a hydrogen donor
by water addition to bio-oil. It was not dried but fed to the hydro- solvent, which may be responsible for much of the reaction, and
treating reactor as a fluid containing 18.0% moisture and having also because the pressure range evaluated was low, from 1.5 to
a density of 1.6 and a moisture-free oxygen content of 20.8% 3.0 MPa of cold hydrogen pressure. Some additional results are
(by difference). The catalyst used was an unspecified Katalco given in a second paper,68 but important information is omitted,
sulfided CoMo pellet. Two tests in a continuous upflow fixed- such as the ratio of bio-oil-to-solvent fed to the reactor, the actual
bed reactor were performed, whose conditions and results are operating pressure, the definition of the “conversion” numbers
presented in Table 23. Comparing the two tests showed little given, or the reasoning for concluding that the optimum
difference in increasing the space velocity from 0.5 to 1; how- conditions are 30 min at 360 °C with a starting cold pressure
ever, the outlet temperature was also higher. Hydrogen con- of 2 MPa, since much higher oil yield was found at 45 min.
sumption was measured at 813 m3/t of liquid feed (3900 scf/
bbl). The light organics product had a H/C ratio of 1.50 with Process Economics
0.46% oxygen by difference. 13C NMR analysis showed a 62/
38 split of aliphatic/aromatic carbon with little evidence of oxygen- Process economics for these biomass liquefaction and bio-
ated components. Simulated distillation by GC showed 65% or oil upgrading processes have been developed by several groups
more with a boiling point below 345 °C and as much as 50% over the period of development in an attempt to determine the
in the gasoline range (below 225 °C). The GC-MS analysis value for proceeding with the research and to focus the research
showed the expected collection of cyclic alkanes and aromatics. effort on the technical issues related to the cost barriers.
East China University. The settled oil phase from the pyrol- Biomass Liquefaction Test Facility (BLTF). The initial
ysis of sawdust was hydrotreated in a batch reactor over a phase of the International Energy Agency Bioenergy (IEA) coop-
sulfided CoMo catalyst. The fast pyrolysis oil was induced to
separate into two phases by water addition. The bottom phase, (67) Zhang, S.-P.; Yan, Y.-J.; Ren, Z.; Li, T. Study of Hydrodeoxygen-
ation of Bio-Oil from the GFast Pyrolysis of Biomass. Energy Sources 2003,
25, 57-65.
(66) Piskorz, J.; Majerski, P.; Radlein, D.; Scott, D. S. Conversion of (68) Zhang, S.-P.; Yan, Y.-J.; Li, T.; Ren, Z. Upgrading of Liquid Fuel
Lignins to Hydrocarbon Fuels. Energy Fuels 1989, 3, 723-726. from the Pyrolysis of Biomass. Bioresour. Technol. 2005, 96, 545-550.
DeVelopments in Hydroprocessing Bio-oils Energy & Fuels, Vol. 21, No. 3, 2007 1815

Table 25. Summary of Capital Costs for Biomass Liquefaction Table 26. Potential for Production and Prices of Various
Processes (in millions of US dollars) Biomass-Derived Liquid Fuels
AFP LIPS energy price EEC potential
t(dry)/ha GJ/t toe/ha ECU/toe Mtoe/y
present potential present potential
ethanol 5-6 26.8 3.3-4.0 530 92
Fixed Capital Investment (FCI)a
biodiesel 1.2-1.6 37.7 1.13-1.51 400-440 33
primary liquefaction 38.9 20.6 65.8 37.8
refined oil 7.5-12.5 40 7.5-12.5 335 ∼245
crude upgrading 36.4 26.8 20.9 20.3
stabilized oil 11.5-17.5 28 8-12.2 260 ∼280
product finishing 11.4 0.5 11.9 0.6
lead-free 230
total 86.7 47.9 98.6 58.7
gasoline
a FCI ) the total capital investment (including major equipment and

instrumentation) plus factors for direct costs, building, site improvement, end of biomass feedstock costs, <$10/t. The upgraded and
utilities, and an indirect cost factor covering engineering and construction refined products were not competitive with existing refined
overheads.
petroleum product costs based on $18.9/bbl petroleum crude
erative project on a biomass liquefaction test facility (BLTF) pre- without tax or other incentives. The production cost of refined
pared technoeconomic assessments of the two biomass liquefac- transportation fuel was determined to be $340/t of oil equivalent
tion process concepts, atmospheric flash pyrolysis (AFP) and (toe) at $35/dry t wood feed, while the market price at the time
liquefaction in pressurized solvent (LIPS) of wood and peat feed- was $185/t. Clearly, these cost numbers need to be re-evaluated
stocks.69 Only initial upgrading results were available at the time, considering the current market environment.
and those results suggested the relatively straightforward hydro- Université Catholique de Louvain. Some important points
treating of the high-pressure oil and the difficulty in applying have been made relative to the perception of the high cost of
conventional hydrotreating to fast pyrolysis oils. The calculated hydrotreating bio-oil.73 The view of the UCL authors was that
product costs for the crude bio-oil, on an energy basis, were low- the hydrotreating of bio-oils is not so different from petroleum
er for the pyrolysis process compared to the high-pressure pro- hydrotreating, is really just an extension of that technology, and
cess. The study also concluded that the process advantages for is probably easier to develop than refining of heavy crudes,
fast pyrolysis included lower investment costs and lower sensi- which was under development at that time (1996). Those authors
tivity of product cost to changes in design parameters and chan- made the point that the amount of hydrogen required for bio-
ges in yield structure. Incorporation of the costs for upgrading oil upgrading is quite comparable to the petroleum industry and
was not attempted based on the early stage of process develop- is less than the requirements for some difficult to hydrocrack
ment. feedstocks. Since those technologies remain profitable, it was
Direct Biomass Liquefaction (DBL). The second phase of concluded that the amount of hydrogen required is not a real
the IEA effort, the direct biomass liquefaction project (DBL), problem. Another objection addressed was the size of the
developed complete process models from biomass to fuel hydrogen plant. Operation at the small scale for bio-oil
products based on either fast pyrolysis or high-pressure liquefac- upgrading was available as a new, but conventional, technology,
tion including hydrotreatment and final refining to liquid fuels.70 and several examples were given. They concluded with the
Estimated capital costs are given in Table 25 for the present important point that upgraded bio-oil (refined fuel or stabilized
case based on existing process data and the potential case based heavy oil) was the least expensive of the biomass-based liquid
on projected improvements in the processes. The study group fuel options and had the most potential to displace petroleum
concluded that there was a slight advantage to fast pyrolysis requirements in Europe (EEC), as shown in Table 26.
over high-pressure liquefaction in that the total energy efficien-
cies were nearly the same (53% for AFP versus 49% for LIPS), Summary
but AFP was determined to be closer to economic feasibility
with a perception of a lesser amount of needed development.71 This review documents the extensive work which has been
On the basis of “potential” case scenarios, it was determined undertaken over the past 25 years in the field of hydroprocessing
that both process options had potential for cost reductions from of biomass-derived liquids. The work extends from small batch-
20 to 30% through process research and development (R&D) reactor tests in universities to demonstration-scale processing
efforts identified in the study. in petroleum-refining laboratories. The potential for the process-
The results of this study were further evaluated using models ing is defined based on these results and is shown to be relatively
developed at Aston University.72 That study concluded that only more expensive than conventional petroleum processing at
the crude bio-oil could be economically competitive with fuel traditional prices as shown by the most recent economic studies,
oil from petroleum in the existing market and only at the low now 10 years old.73 The same study found that pyrolysis-derived
oils are more economical than other biomass-to-liquid fuel
(69) McKeough, P.; Nissilä, M.; Solantausta, Y.; Beckman, D.; Östman, processing, such as ethanol or biodiesel. The potential for
A.; Bergholm, A.; Kannel, A. Techno-Economic Assessment of Selected competitive costs in the fuels marketplace with the recent
Biomass Liquefaction Processes; IEA cooperative project D1, Biomass increases in petroleum prices needs to be evaluated.
Liquefaction Test Facility Project final report, Vol. 5, National Technical
Information Service (DOE/NBM-1062): Springfield, VA, 1988.
(70) Beckman, D.; Elliott, D. C.; Gevert, B. S.; Hörnell, C.; Kjellström, Acknowledgment. The support for preparation for publication
B.; Östman, A.; Solantausta, Y.; Tulenheimo, V. Techno-Economic As- of this review was provided by the Office of the Biomass Program
sessment of Selected Biomass Liquefaction Processes; report of the IEA
Direct Biomass Liquefaction activity, VTT Report 697, VTT: Espoo,
of the U.S. Department of Energy. The Pacific Northwest National
Finalnd, 1990. Laboratory is operated for the Department of Energy by Battelle
(71) Elliott, D. C.; Baker, E. G.; Beckman, D.; Solantausta, Y.; under contract AC0676RLO1830.
Tolenhiemo, V.; Gevert, S. B.; Hörnell, C.; Östman, A.; Kjellström, B.
Technoeconomic Assessment of Direct Biomass Liquefaction to Transporta- EF070044U
tion Fuels. Biomass 1990, 22, 251-269.
(72) Cottam, M.-L.; Bridgwater, A. V. Techno-Economics of Pyrolysis (73) Grange, P.; Laurent, E.; Maggi, R.; Centeno, A.; Delmon, B.
Oil Production and Upgrading. AdVances in Thermochemical Biomass Hydrotreatment of Pyrolysis Oils from Biomass: Reactivity of the Various
ConVersion; Bridgwater, A. V., Ed.; Blackie Academic & Professional: Categories of Oxygenated Compounds and Preliminary Techno-Economical
London, 1993; Vol. 2, pp 1343-1357. Study. Catal. Today 1996, 29, 297-301.

You might also like