Sie sind auf Seite 1von 44

Catalysis Today 217 (2013) 1356

Contents lists available at ScienceDirect

Catalysis Today
journal homepage: www.elsevier.com/locate/cattod

Review

Hydroprocessing challenges in biofuels production


Edward Furimsky
IMAF Group, 184 Marlborough Avenue, Ottawa, Ontario, Canada K1A 8G4

a r t i c l e

i n f o

Article history:
Received 2 November 2012
Accepted 23 November 2012
Available online 12 February 2013
Keywords:
Biomass conversion
Biofeeds hydroprocessing
Biofuels production
Biodiesel properties
Hydroprocessing catalysts
Coprocessing biofeeds with petroleum
feeds

a b s t r a c t
Biofuels production from biomass of lignocellulosic, vegetable oils and algae origins as well as from
municipal solid waste via hydroprocessing (HPR) is in various stages of development. The conversion
of biomass from these sources to biofeeds and chemical composition of the latter are presented. Differences between the mechanism and kinetics of HPR reactions occurring during the HPR of biofeeds
and petroleum feeds are evaluated. Fundamental aspects of conventional and non-conventional HPR
catalysts, with emphasis on their applications in biofuels production are discussed. Catalysts exhibiting
high activity and stability under conditions encountered during the HPR of biofeeds are identied. They
include catalysts consisting of conventional metals (Mo/W and Co/Ni) supported on various supports
as well as novel catalytic phases containing noble metals as well as phosphides, carbides, nitrides and
borides of transition metals in combination with supports varying widely in surface acidity. The studies
on coprocessing biofeeds with the feeds of petroleum origin as well as those on blending biofuels with
petroleum fuels were reviewed. Improvement in properties of petroleum diesel, particularly in terms of
diesel number, can be achieved by blending with biodiesel from vegetable oil sources.
Developments in upgrading biofeeds in aqueous environment (subcritical water, supercritical water
and supercritical alcohols) in the presence of various catalysts and hydrogen, are addressed.
Crown Copyright 2013 Published by Elsevier B.V. All rights reserved.

Contents
1.
2.

3.

4.

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Conversion of biomass to biofeeds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.1.
Conversion of vegetable oils to biofeeds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.2.
Conversion of lignocellulosic biomass to biofeeds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.2.1.
Biomass liquefaction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.2.2.
Biomass pyrolysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.3.
Conversion of algae to biofeeds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.4.
Conversion of sewage sludge to biofeeds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.5.
Biofeeds from gas-to-liquids conversion of biomass . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Composition of biofeeds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.1.
Vegetable oil origin biofeeds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.2.
Lignocellulosic biofeeds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.3.
Biofeeds of algae origin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.4.
Biofeeds from sewage sludge . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.5.
Biofeeds from FTS processes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Catalysts for hydroprocessing of biofeeds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.1.
Conventional HPR catalysts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.1.1.
Modication of active phase during HPR of biofeeds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.2.
Non-conventional catalysts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.2.1.
Noble metals containing catalysts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.2.2.
Metal carbides, nitrides and phosphides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.3.
Role of support . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

E-mail address: edfurimsky@hotmail.com


0920-5861/$ see front matter. Crown Copyright 2013 Published by Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.cattod.2012.11.008

14
15
15
16
16
16
16
19
19
19
19
19
20
22
22
22
22
24
26
26
27
27

14

E. Furimsky / Catalysis Today 217 (2013) 1356

5.

Reactions occurring during HPR of biofeeds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .


5.1.
Mechanism of HPR reactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.1.1.
Reactants of ligno-cellulosic origin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.1.2.
Reactants in vegetable biofeeds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.1.3.
Reactants in algae biofeeds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.1.4.
FTS reactants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.2.
Kinetics of HPR of bio-feeds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.2.1.
Model compounds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.2.2.
Real bio-feeds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
6.
Catalyst development for HPR of biofeeds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
6.1.
Catalysts for HPR of biofeeds of ligno-cellulosic origin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
6.1.1.
Conventional catalyst . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
6.1.2.
Unconventional catalysts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
6.2.
Catalysts for HPR of vegetable oils biofeeds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
6.2.1.
Conventional HPR catalysts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
6.2.2.
Non-conventional catalysts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
6.3.
Catalysts for HPR of algae biofeeds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
7.
Coprocessing of biofeeds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
8.
Hydroprocessing of biofeeds in aqueous phase . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
8.1.
Sub-critical conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
8.2.
Super critical conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
9.
Hydroprocessing of biofeeds in protic solvents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
10.
Renery aspects of biofuels production . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
11.
Observations and future perspectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

List of acronyms
CCR
Conradson carbon residue
coal derived liquids
CDL
CFPP
cold lter plugging point
CUS
coordinatively unsaturated sites
DBF
dibensofuran
DBT
dibenzothiophene
DDO
direct deoxygenation
DDS
direct desulfurization
DMDBT dimethyldibenzothiophene
DMDS dimethyldisulde
DMS
dimethyl sulde
DFT
density functional theory
EN
European norm
fatty acids methyl esters
FAME
FBP
nal boiling point
uid catalytic cracking
FCC
FTS
FischerTropsch synthesis
GCMS gas chromatographymass spectroscopy
GTL
gas-to-liquids
HAAD-STEM high angle annular dark-eld scanning transmission electron microscopy
HCR
hydrocracking
HDI
hydroisomerization
hydrodenitrogenation
HDN
HDO
hydrodeoxygenation
HDS
hydrodesulfurization
HPR
hydroprocessing
HRTEM high resolution transmission electron microscopy
HYD
hydrogenation
LC
liquid chromatography
LCO
light cycle oil
LH
LangmuirHinshelwood
LHSV
liquid hourly space velocity
NMR
nuclear magnetic resonance

SCW
STM
THF
TPD
TPS
ULSD
USDA
VGO
WCO
WGS
WHSV
XPS

28
28
28
33
34
34
35
35
37
39
39
39
40
42
42
45
48
48
49
49
50
51
51
52
53

supercritical water
scanning tunneling microscopy
tetrahydrofuran
temperature programmed desorption
temperature programmed sulding
ultra low sulfur diesel
US department of agriculture
vacuum gas oil
waste cooking oil
water gas shift
weight hourly space velocity
X-ray photoelectron spectroscopy

1. Introduction
In the comprehensive reviews on different aspects of biofuels
published by Corma et al. [1,2], Naik et al. [3] and Demirbas [4],
current methods and future opportunities for the production of
transportation fuels from various sources of biomass, were presented. This included the chemistry and catalysis of processes,
biorenery concepts, engineering challenges and solutions. In the
case of economic analyses, a life cycle which begins with biomass
planting, growing, harvesting and biomass preparation for the production of biofeeds, was not always considered. The authors [14]
pointed out that economic analyses are very site specic and may
vary considerably depending on the current and future regional
economic and market situations.
There has been a tendency to classify biofuels into three generations to reect the level of development and commercialization
[5,6]. The rst generation biofuels such as ethanol and biodiesel
have already been commercially available. The latter is represented
by fatty acids methyl esters (FAME). Detailed accounts of the FAME
production and properties were given in the comprehensive review
published by Van Gerpen [7]. The rst generation biofuels are of
the food-based biomass origin. In this regard, corn, potatoes, beet,
sugarcane, etc. are primary sources of commercial ethanol, while
vegetable oils derived from rapeseed, olive, sunower, soya, palm,

E. Furimsky / Catalysis Today 217 (2013) 1356

etc. are the feeds for production of FAME. The cetane number of
FAME is in the range of that of commercial diesel (53), but is
much lower than that of the diesel derived from vegetable oils via
hydroprocessing (HPR), e.g., 8099 [8]. Because of good lubricating
properties, the presence of FAME in diesel fuel reduces fuel system wear and increases life of the fuel injection system. However,
storage stability of the FAME based biodiesel requires attention.
It has been suggested that the second generation biofuel such
as ethanol made from inedible cellulosic plant matter (e.g., agricultural and forest residues and fast growing crops such as coppices)
may ensure sustainability of biofuel production [5,6]. In this case,
an enzyme aided biochemical conversion of lignocellulosic structures to sugar-like structures is necessary to obtain bio-ethanol by
fermentation. For this route, signicant advances in the research
on enzyme involvement in these reactions are still needed for the
second generation ethanol to have an impact on fuel supply.
For more than three decades, the biofuels produced from lignocellulosic sources (e.g., straw, switchgrass, sawdust, wood chips,
etc.) via either pyrolysis or hydrothermal liquefaction, have been
receiving much attention. The extensive database in the literature
conrmed that the developments in this eld reached a near commercial stage [14]. Yet, this source of biofuels was not clearly
identied among the above categories. From this source, the liquid
fuels meeting specications of commercial fuels can be prepared
via HPR route, although this may require signicant amounts of H2
and catalyst.
Microalgae have been identied as a potential source of the
third generation biofuels because of their higher photosynthetic
efciency, faster growth rate, and higher area-specic yield compared with other sources of biomass. Thus, annually, microalgae
can produce up to 100,000 L of oil per hectare compared with about
5950, 2689, 1413 and 952 for palm, coconut, castor and sunower,
respectively [9].
The gas-to-liquid (GTL) is another route for the production of
biofuels. It involves thermochemical conversion of biomass to synthesis gas (syngas) using either pyrolysis or gasication. The syngas
can then be converted to fuels via FischerTropsch synthesis (FTS).
While the FTS process has been commercially available for decades,
some efciency and environmental issues associated with syngas
production from biomass still need to be addressed.
In recent years, growing interests in the development of green
fuels (diesel and gasoline) from vegetable oils has been noted. In
this case, rather than being converted to FAME via transesterication, fatty acids are completely converted to hydrocarbons via
HPR [10,11]. If an appropriate HPR catalyst and operating conditions are chosen, the nal product may be used either directly as
green diesel or blended with diesel fractions of petroleum origin.
Besides much higher cetane number, the advantage of green diesel
compared with the FAME containing diesel is a higher stability of
the former. Moreover, cold ow properties (pour point, cloud point
and freezing point) specications limit the amount of FAME to be
blended with the base diesel. Contrary to this, cold ow properties
of the green diesel may be adjusted during HPR over bifunctional
catalyst with a high hydroisomerization (HIS) activity and
selectivity.
The properties of biodiesel (FAME), green diesel, FTS diesel
and commercial EN 590 standard diesel are shown in Table 1
[3]. High cetane number of green diesel and FTS diesel should be
noted. Similarly, a high quality bio gasoline may be prepared from
biofeeds. Thus, in the case of bio gasoline derived from vegetable oil
sources, a high octane number may be achieved by selecting a catalyst exhibiting a activity for hydrocrackinghydroisomerization
(HCRHIS).
The HPR in sub-critical and super-critical water environment as
well as in supercritical alcohols have been attracting attention. The
novel processes are particularly suitable for the HPR of biofeeds.

15

Table 1
Comparison of properties of FAME and green diesel.
Properties

FAME

Green diesel

GTL diesel

EN 590

Cetane number
Density (g/mL)
NOx emissions (%)
Sulfur (ppm)
Oxygen (wt.%)
Total aromatics (wt.%)
Boiling range ( C)
Cloud point ( C)
Storage stability

50
0.883
+10
<10
11
0
340355
5
Challenging

8090
0.78
0 to 10
<10
0
0
265320
5 to 30
Good

7381

53

<10
0
0
190330
0 to 25
Good

<10
0
30
180360
5
Good

Ref. [3].

The latter chapters of this review provide summary of the most


recent developments in this new eld.
2. Conversion of biomass to biofeeds
The HPR method may be used for production of commercial
fuels from biofeeds obtained from vegetable oils, lignocellulosic
biomass, algae biomass and activated sewage sludge. Therefore,
biofeeds are in fact biocrudes and/or partially upgraded biocrudes,
which require additional upgrading via HPR to commercial fuels.
The synthetic crude from FTS has been referred to as biofeed in
the case, that syngas was obtained from biomass. It has been evident that processing conditions, catalyst inventory and hydrogen
consumption vary widely depending on the origin of biofeeds [14].
2.1. Conversion of vegetable oils to biofeeds
At room temperature, vegetable fats and animal fats as well
as vegetable oils derived from plants are solids and liquids,
respectively. In their composition, triglycerides are predominant
structures, however various concentrations of free fatty acids,
monoglycerides and diglycerides as well as unsaponiable lipids
may also be present. Both edible and inedible vegetable oils and fats
can be found. The former include vegetable oils from soy, canola,
sunower, safower, peanut, cottonseed, etc., while linseed oil,
tung oil and castor oil are typical inedible oils.
According to the USDA, in 2007/2008, the total world consumption of major triglyceride vegetable oils exceeded 130 million tons.
Table 2 [12] shows the contribution of the different sources of vegetable oils to the total amount consumed.
Vegetable oils are recovered from the plant seeds and other
parts of plant using chemical extraction, mechanical extraction
and expeller-pressing extraction. Chemical extraction employs
Table 2
Total world consumption (million tons) of vegetable oils in 2007/2008.
Origin

Consumption

Selected application

Palm
Soybean

41.3
41.3

Rapeseed

18.2

Biofuel production
About half of
worldwide edible oil
production
Widely used cooking
oils (canola)
Cooking oil, also used
to make biodiesel
Cooking oil
Industrial food
processing

Sunower seed

9.9

Peanut
Cottonseed

4.8
5.0

Palm kernel
Coconut
Olive

4.9
3.5
2.8

Ref. [12].

Soaps and cooking


Cooking, cosmetics,
soaps and fuel for oil
lamps

16

E. Furimsky / Catalysis Today 217 (2013) 1356

Table 3
Property ranges of bio-crude obtained by liquefaction and pyrolysis.

Carbon (wt.%)
Sulfur + nitrogen (wt.%)
Oxygena (wt.%)
Water in crude (wt.%)
Density (g/cm3 )

Liquefaction

Pyrolysis

6881
0.1
925
625
1.101.14

5666
0.1
2738
2452
1.111.23

Ref. [18].
a
Dry basis.

parafnic solvents and gives the highest yield of oil compared


with the other methods. Supercritical CO2 extraction may also
be used [13]. The efciency of extraction always depends on the
solubility of vegetable oil in the extracting solvent. Different types
of oil presses have been used for the expeller-pressed extraction.
However, the oil yield was always the lowest among the three
extraction methods. Crushed seeds are usually used for mechanical
extraction. Chemical extraction would be the most suitable method
in the case that vegetable oils are considered as the biofeeds for
production of biofuels.
Novel processes for production of biofeeds from vegetable oil
and animal fat sources include hydrothermal liquefaction and
pyrolysis. The hydrothermal liquefaction process described by Li
et al. [14] operated at a high pressure (e.g., 17 MPa) and temperatures ranging from 450 to 475 C, i.e., in a super critical water (SCW)
region with and without a catalyst. Under these conditions, water
serves as a reactant, catalyst, and solvent for typically acid- or basecatalyzed reactions. Apparently, water may also supply hydrogen
for cracking, hydrolysis of triglycerides, followed by decarboxylation. The high-temperature and high-pressure water reduces the
formation of gaseous products and minimizes the formation of
chars. For vegetable oils containing more than 30% polyunsaturated
fatty acids, the products had a high content of aromatic structures.
The pyrolysis of animal fats and vegetable oils may also be
used to produce biofeeds, although the information on this route
of biofeeds production is limited. This may be, at least partly
attributed to the loss of valuable liquid hydrocarbons to gas and
coke, generally observed during pyrolysis of biomass. The studies
conducted by Simacek et al. [15,16] may be used to illustrate further upgrading of the biofeeds derived from vegetable oils to green
diesel.
2.2. Conversion of lignocellulosic biomass to biofeeds
The rst report on potential of lignocellulosic biomass as the
source of liquid fuels was published by Heinemann [17] in 1954.
Two groups of processes, i.e., high pressure (under N2 or H2 O)
hydrothermal processes with relatively long residence time as
well as pyrolysis processes employing either a short contact time
(fast-pyrolysis) or a long contact time (slow-pyrolysis), have been
evaluated [18,19]. Detailed accounts of hydrothermal methods
used for biomass conversion, were given by Peterson et al. [20]
and Demirbas [21]. The ranges of elemental composition, water
content and density of the primary products from these processes
are shown in Table 3 [18]. Generally, the content of oxygen and
water in the bio-crude from pyrolysis is greater than that from
hydrothermal liquefaction, whereas the yield of bio-crude from the
latter is greater than that from pyrolysis. Therefore, the HPR of the
pyrolysis biocrude is more challenging [19,22].
2.2.1. Biomass liquefaction
Zhang et al. [23] reported that heating time inuenced yields of
liquid products. In some studies [24,25], the bio-crude was obtained
during the liquefaction of wood chips at 623 K, 18 MPa and 6 min
contact time using the solution of a caustic, i.e., presumable Na2 CO3 ,

in water. Under such conditions, a signicant decarboxylation of


biomass was observed. This was evidenced from high yields of
carbon dioxide.
The liquefaction of lignocellulosic biomass (e.g., wood meal
birch and aspen, jute bers, thermomechanical pulp, kraft pulp,
cotton and kenaf plant meal) was conducted hydrothermally and
in the presence of phenol [26]. The conversions in compressed
water (10 MPa) were much lower compared with those in phenol.
The dissolution of biomass increased with increasing temperature and reached an optimum at 250 C and 45 min reaction time.
The concentration of catalyst (Na2 CO3 ), the phenol/biomass ratio,
type of biomass and pH were additional parameters inuencing
the overall conversion, the quality and yield of the liquid products. In another mixed solvent biomass liquefaction study [27], the
50/50 mixture of either methanol/water or ethanol/water were the
most effective during the liquefaction of a sawdust at 300 C and
15 min reaction time. Under these conditions, the yield of bio-oil
approached 65 wt.% at the overall biomass conversion of greater
than 95%. Above 300 C, an excessive conversion of bio-oil to char
became evident. A depolymerization of biomass was observed at
330 C and 18 MPa (410 min) even without catalyst and/or an
agent being present [28]. A high level of biomass depolymerization was achieved during liquefaction conducted in hot compressed
water at 503 K (2.8 MPa) over the Ru catalyst supported on mesoporous carbon [29].
The similarity of conditions employed during coal liquefaction
was explored during the co-processing of agricultural or biomass
waste (sawdust, horse manure, cow manure, super manure) with
coal [30]. An increase in the conversion of coal by the coliquefaction was observed. Matsumura et al. [31] carried out the
co-liquefaction of biomass and coal under SCW conditions. Akash
et al. [32a] carried out co-liquefaction of coal and lignin under pressurized H2 in tetralin. They reported that the oil yield during the
co-liquefaction increased. Co-liquefaction products contained large
amount of low boiling compounds.
The two-step process comprising liquefaction in an acidic subcritical water followed by the supercritical liquefaction in the
Ca(OH)2 containing water was used for conversion of the switchgrass biomass to biocrude [32b]. Hemicelluloses were liqueed in
the rst step carried out at 200 C to minimize the formation of char.
The remaining biomass (mainly polysaccharides) was liqueed in
the second stage carried out at 380 C. Compared with the one step
process, the yield of biocrude increased using the two steps process.
This may be attributed to diminishing of the char forming reactions
in the latter process.
2.2.2. Biomass pyrolysis
Detailed accounts of pyrolysis methods used for conversion of
biomass to liquids were given by Mohan et al. [33a] and Bridgwater [33b]. Because of high content of guaiacols and phenols,
liquids from pyrolysis are less stable than those from liquefaction
[3438a]. The quality of pyrolysis liquids can be improved by an
in situ upgrading with the aid of catalyst as reported by Iliopoulou
et al. [38b].
2.3. Conversion of algae to biofeeds
Among different aquatic sources (oceans, lakes, estuaries, etc.),
the waste waters derived from municipal, agricultural and industrial activities were identied as additional sustainable means
of algal growth [9,38c,39]. An opportunity exists for combining
wastewater treatment by algae via nutrient removal with biofuel production. The experimental systems and conditions used
to study algae cultivation include reactors employing horizontal
tubes, vertical tubes and at plates [40a]. Other stages, i.e., harvesting, dewatering, lipids extraction and conversion to biofeeds

E. Furimsky / Catalysis Today 217 (2013) 1356


Table 4
Biochemical composition (wt.%) of microalgae strains (daf).

Table 6
Elemental analyses of microalgae and biooils from liquefaction and pyrolysis.

Microalgae strain

Protein

Carbohydrate

Lipid

Chlorella vulgaris
Nannochloropsis oculata
Porphyridium cruentum
Spirulina

55
57
43
65

9
8
40
20

25
32
8
5

Ref. [41].

Table 5
Elemental analysis of microalgae.

C
H
H/C
O(diff)
N
S
Protein
Fat
Fatty acid
Carbohydrate
Ash

17

Analysis (wt.%)

Carbon
Hydrogen
Nitrogen
Sulfur
Oxygendiff
H/C ratio
O/C ratio

Algae

46.2
7.1
10.6
0.74
35.4
1.84
0.57

Liquef.

Pyrolysis

350 C

350 C

500 C

73.7
8.9
6.3
0.9
10.2
1.44
0.10

67.5
9.8
10.7
0.45
11.3
1.73
0.13

74.7
10.6
7.1
0.81
6.8
1.68
0.06

Ref. [44].

Spirulina

Chlorella

Littorale

46.1
7.4
1.92
41.4
4.8
0.4
57.5
12.0
1.0
<0.5

47.3
7.2
1.83
37.6
8.2
0.7
80.0
10.0
0.8
<0.5
0.2

35.5
5.4
1.82
53.1
6.0

37.6
9.9
6.4
23.0
29.5

Ref. [41].

were described by Knoshaug at al. [40b]. Advancement in these processes, from laboratory scale to a commercial production received
attention as well [40c].
Amin [41] described biochemical (e.g., fermentation and anaerobic digestion) and thermochemical (pyrolysis and liquefaction)
methods for conversion of algae to biofeeds. The former processes
produce the rst generation biofuels such as ethanol and biodiesel,
whereas thermochemical processes produce biofeed and gas. The
solvent extraction of biofeed was successfully applied to microalgae as well [42]. Among different plant oils, microalgae have the
highest oil yield. For example, on a dry basis, the oil content of
some microalgae (e.g., Botryococcus braunii) may approach 80%. A
brief account of the methods used for production and harvesting
microalgae was included in the review published by Amin [41] as
well. The information on the growth and culture of micro-algae,
which have a potential for the liquid fuel production was reviewed
by Tran et al. [43a]. In this review, chemical composition of algae
was also discussed.
Approximate biochemical compositions of several microalgae,
expressed on dry-ash-free (daf) basis are shown in Table 4 [41]. In
every case, proteins are the main component in addition to variable
contents of peptides and carbohydrates. Because of different structures of these components, variable product distribution among
the algae of different origin may be anticipated. For example, a
high protein containing algae should have a high nitrogen content. Table 5 [41] shows the analyses of dried microalgae such as
Chlorella, Spirulina, and Littorale. Drying decreased moisture content from more than 80% to less than 30%. Thus, if the results for
Chlorella and Spirulina are expressed on a daf basis, they should
approach those in Table 4. Relatively high hydrogen content and
H/C ratio should be noted. However, a portion of this hydrogen is

Fig. 1. Aminoacids building block of proteins.

associated with O H and N H bonds rather than C H bonds as


evidenced by the tentative structure of protein molecules in Fig. 1.
The direct extraction of bio-oil from algae was investigated
by Kanda et al. [43b]. In their study, the paste prepared from
B. braunii Race B was extracted in an extraction column (ID of
11.6 mm and length of 190 mm) at room temperature using DME
solvent. The extracts contained large amount of C32 C34 botryococcenes. This method was compared with hexane Soxhlet extraction.
In this case, the B. braunii cells were mechanically homogenized.
Although the overall yields were similar (49 wt.% on dry basis), the
hexane extract contained small amounts of C30 and C31 botryococcene in addition to C32 C34 . It also had a lower nitrogen a higher
hydrogen content than the DME extract.
Jena and Das [44] compared slow pyrolysis method with hydrothermal liquefaction for production of biooils from microalgae. In
nonisothermal pyrolysis mode at 350 and 500 C, heating of the
sample was carried out for 3.5 and 7 C/min, respectively. In an
isothermal mode, the pyrolysis was conducted for 60 min. For liquefaction experiments, 150 g of algae (daf) was slurried with 600 g
of deionized water in a 1.8-L stirred reactor. After purging and
pressurizing with nitrogen to 2 MPa, the reactor was heated to
350 C attaining the water pressure of about 20 MPa and holding for
60 min. The elemental analyses of biooils are compared in Table 6
[44] with that of algae biomass. The yield of bio-oil from liquefaction approached 41% compared 2329% for pyrolysis. Based on the
content of nitrogen and oxygen, the biooil quality from liquefaction
was much higher than that produced by pyrolysis at 350 C. However, pyrolysis at 500 C produced biooil in its quality comparable
to the liquefaction product.
The hydrothermal liquefaction of microalgae Desmodesmus sp.
was investigated at temperatures of 175450 C and reaction times
up to 60 min in a batch reactor system [45a,b]. A major rupture of
the cells by increasing temperature from 225 to 250 C at 5 min
contact time, was noted. The highest yield of oil (49 wt.%) was
obtained at 375 C and 5 min contact time. This represented about
75% of the algal energy.
Ross et al. [46] investigated the effect of operating conditions on
quality of liquid products from microalgae and cyanobacteria such
as Chlorella vulgaris and Spirulina. The catalysts included potassium
hydroxide and sodium carbonate as well as organic acids such as
acetic acid and formic acid. The organic acids catalysts gave higher
yields of bio-crude. This biocrude had a lower boiling point, better cold ow properties, carbon content of 7075% and an oxygen
content of 1016%. The nitrogen content in the bio-crude produced during the alkali catalyzed liquefaction ranged from 4 to
6 wt.%, while in the presence of organic acids the nitrogen content
approached 7 wt.%. Nitrogen distribution (aqueous phase, biocrude,
residue and gaseous phase) could be inuenced by the type of catalyst, origin of algae and experimental conditions. For example,
the amount of nitrogen in aqueous phase decreased with increasing temperature. The amount of nitrogen in gaseous phase was

18

E. Furimsky / Catalysis Today 217 (2013) 1356

Fig. 2. Schematics of extractions of algae liquefaction products [48].

inuenced by the type of catalysts and increased in the following


order: CHOOH > KOH > Na2 CO3 > CH3 COOH [46].
Algal cells of Dunaliella tertiolecta with a moisture content of
78.4 wt.% and dry solid content of 21.6 wt.% was thermochemically liqueed in a batch reactor using 20 g of algae (5 g of dry
solid) between 200 and 340 C [47,48]. Before the experiments, the
reactor was purged with N2 to displace air and pressurized by N2
to 3 MPa to prevent evaporation of water. The reaction mixture
was extracted using CH2 Cl2 according to the procedure shown in
Fig. 2 [47,48]. Depending on experimental conditions (temperature,
duration and with or without Na2 CO3 catalyst), the oil yield on an
organic basis varied between 31 and 44 wt.%. For selected runs, elemental analysis of oils is shown in Table 7 [47]. It is evident that
the oxygen and nitrogen contents decreased and that of carbon
and hydrogen increased during liquefaction. For the best run (#5),
almost 80% of oxygen was removed compared with only about 30%
of nitrogen removal.
Dote et al. [48,49] performed hydrothermal liquefaction of B.
braunii microalgae according to procedure used by Minowa et al.
[47]. Hydrocarbons separation from the liquefaction products was
carried out according to the procedure shown in Fig. 2 as well as
via hexane extraction. During the latter method, the freeze dried
algae cells were sonicated with hexane for 30 min at 30 C. The
suspension was then ltered and hexane evaporated.
From the HPR point of view, it is essential that during hydrothermal liquefaction, most of the algaes nitrogen ends up in
aqueous phase rather than in the bio-oil phase. In this regard,
promising results were published by Valdez et al. [50] who

determined the distribution of carbon, hydrogen and nitrogen


among bio-oil and aqueous phases during the hydrothermal liquefaction of Nannochloropsis sp. microalgae at 350 C for 60 min.
Most of the carbon and hydrogen in the algal biomass ended up in
the bio-oil product while most of the nitrogen appeared in aqueous phase in the form of ammonia. However, in spite of this, the
crude bio-oil still contained about 4 wt.% of nitrogen. The recovery
of bio-oil from the product mixture was inuenced by the type of
solvent. For example, for hydrocarbon solvents such as hexadecane
and decane, the yield of bio-oil approached 39 wt.%. The carbon content of this bio-oil was about 69% compared with about 75% for the
bio-oil obtained with chloroform and dichloromethane solvents.
Bio-oil recovered with the latter solvents had a higher content of
free fatty acids compared with that obtained with hydrocarbon solvents. Based on the results published by Zhou et al. [51], there was
little advantage in performing the liquefaction of marine macroalgae Enteromorpha prolifera (one of the main algae genera for green
tide) at 300340 C in the presence of catalyst (Na2 CO3 ). Thus, the
differences between the results obtained with and without the catalyst were within the accumulated error. Similarly, little effect of
the catalyst on the liquefaction yields was reported by Dote et al.
[48].
Matsui et al. [52] studied catalytic and non-catalytic liquefaction of high-protein Spirulina microalgae (300425 C; 5.0 MPa
H2 , batch reactor) in different solvents (e.g., tetraline, methylnaphthalene and toluene) and either in H2 or N2 atmosphere. The
Fe(CO)5 -S was used as catalyst. More than 90% of the reaction product in tetraline was THF soluble. The hexane soluble portion of

Table 7
Yields and properties of oils.
Run

Temp. ( C)

Time (min)

Na2 CO3 (wt.%)

Yield (wt.%)

0
1
2
3
4
5
6
7
8

Ref. [47].
a
Algae feed.

250
250
250
250
340
340
340
340

5
5
60
60
5
5
60
60

0
5
0
5
0
5
0
5

34.3
34.2
33.3
34.2
37.8
35.7
36.2
33.6

Elemental analysis (wt.%)


C

H/C

53.3
72.4
71.5
71.3
71.6
76.9
74.0
75.0
76.3

5.2
8.4
8.6
8.1
8.5
9.1
8.7
8.7
8.4

1.17
1.39
1.44
1.36
1.42
1.52
1.41
1.39
1.32

9.8
6.5
6.9
6.9
7.4
6.6
6.9
6.6
6.9

31.7
12.7
13.0
13.7
12.5
7.4
10.4
9.7
8.4

E. Furimsky / Catalysis Today 217 (2013) 1356

THF solubles varied from 52.3 to 66.9 wt.%. The former portion was
referred to as oil. However, higher yields of oil, i.e., 78.3 wt.% at
350 C, were obtained in the presence of water even under N2 but
oil had a low carbon and high oxygen contents. In toluene solvent,
the oil had a high carbon content and lower oxygen content. Benecial effect of catalyst was conrmed by increased yields of oil, while
that on the overall conversion was less clear. Several catalysts (e.g.,
Pd/C, Pt/C, Ru/C, Ni/SiO2 Al2 O3 , sulded CoMo/Al2 O3 and zeolite)
were used by Duan and Savage [53] for the hydrothermal liquefaction of the microalga Nannochloropsis sp. at 350 C, both under He
and under a high-pressure of H2 . With respect to the yield of bio oil,
the advantages of the presence of catalyst and H2 were not clearly
evident.
The potential of co-liquefaction of coal and microalgae under
coal liquefaction conditions was explored by Ikenaga et al. [54].
The catalysts such as Fe(CO)5 -S, Mo(CO)6 -S, and Ru3 (CO)12 were
slurried with coal-algae (one-to-one ratio) mixture in 1-methyl
naphthalene. Microalgae of Chlorella, Spirulina, and Littorale type
(Table 5) [41] as well as Yallourn and/or Illinois No. 6 coal, were
used. The conversion dened as THF solubles of reaction mixture
approached 99%, whereas the hexane soluble portion of THF solubles dened as oil approached 66 wt.%. Both, the conversion and
the yield of oil approached those obtained by the addition of yields
from the separate runs using either coal or algae alone.
Synergetic effects involving Spirulina microalgae and a highdensity polyethylene during the co-liquefaction in sub- and
supercritical ethanol were reported by Pei et al. [55a]. This was
conrmed by an increased yield and quality of liquid products. For
example, carboxylic acids, esters and ketones were dominant components after liquefaction of Spirulina algae alone, whereas after
the co-liquefaction, the product mixture was dominated by hydrocarbons.
2.4. Conversion of sewage sludge to biofeeds
The activated sewage sludge from waste sewage treatment
plants is potential source of biofeeds. It can be converted to biofeed
by pyrolysis. In the study published by Cao et al. [55b] sewage
sludge was dehydrated and anaerobically digested before being fed
into the internal uidized bed pyrolyser. The pyrolyses performed
at 500 C resulted in 45.2 wt.% yield of biooil. The homogeneity of
sewage sludge was rather low. Therefore, a wide range of composition of the pyrolysis-oils may be obtained. According to Izhar
et al. [55c], the pyrolysis oil may contain between 59 wt.% nitrogen, 5675% carbon, 1324% oxygen, 819% hydrogen and 2% ash.
A high content of nitrogen compared with the biofeeds of vegetable
oils and lignocellulosic origins should be noted. In addition, during upgrading via HPR, a high viscosity of such biofeed may affect
handling due to very low owability/pumpability [55b,c].
Rushdi et al. [55d] performed two sets of experiments between
200 and 350 C with contact time of 48 h. In one case, the reaction mixtures consisted of sewage sludge and water, whereas the
second set also contained oxalic acid as the source of an in situ
hydrogen. The yields increased with aqueous oxalic acid by a factor
of 1.4 compared with oxalic acid free conditions.
2.5. Biofeeds from gas-to-liquids conversion of biomass
Conversion of biomass to syngas using either pyrolysis or gasication is the rst step during biofuel production via gas-to-liquid
(GTL) method. Any kind of biomass can be converted to syngas
(CO + H2 ). After purication, followed by the H2 /CO ratio adjustment in the water-gas-shift (WGS) process, syngas is used as the
feed for FTS process. It is believed, that a lower efciency of synthesis gas production from biomass and higher associated emissions
compared with other sources of syngas, need to be addressed

19

Table 8
Typical elemental compositions of bio-crude from different sources.
Properties

Biocrude
Veget. oil

C (%)
H (%)
H/C
O (%)
N (%)
Density (kg/L)

77.9
11.7
1.8
10.4
0.04
0.89

Wood

Microalgae

Pyrol.

Liquef.

Pyrol.

Liquef.

56.4
6.2
1.3
37.3
<0.1
1.21

64.5
7.7
1.4
22.5
<0.1
1.12

61.5
8.5
1.7
20.2
9.8
1.2

74.0
8.7
1.4
10.4
6.9
1.1

[56]. However, some positive conclusions were made in the study


published recently by Manganaro et al. [57]. In the proposed
scheme, biomass was harvested, locally pyrolyzed to bio oil which
was transported to a remote processing facility for autothermal
reforming to produce syngas for FTS plant. This scheme was evaluated in details with focus on energy balance, efciency and
economic feasibility. Martin and Grossmann [58] compared the GTL
route for the conversion of biomass with the biomass liquefaction
option. An extensive database on technological aspects of FTS and
quality of liquid products was presented in several authoritative
reviews [5962].
3. Composition of biofeeds
The objective of this sub-chapter is to give a summary of the
structural and compositional differences among of various biofeeds
relevant for selection of optimal HPR conditions [38,6365]. The
previous chapter identied several differences based mostly on elemental analysis. As Table 8 shows, structural differences are not
clearly evident from such information. More extensive evaluations
using chromatographic and spectroscopic techniques as well as
some physical testing of biofeeds, are necessary.
3.1. Vegetable oil origin biofeeds
The main fatty acids which are part of glycerides in vegetable oils
and the corresponding biofeeds are listed in Table 9 [14]. An extensive characterization of the biofeeds obtained by hydrothermal
liquefaction of waste lard and locust leafs was conducted by Chen
et al. [64]. The composition of lard was dominated by fatty acids
while that of the locust leafs by lignin cellulose and hemicellulose.
The biofeed obtained by the liquefaction of the mixture of locust
leaves and lard in the weight ratio of 4.4/1 was dominated by the
components of the lignocellulosic origin. This appears to be the
only study in which a mixture of the lignocellulosic biomass and
glycerides was used for biofeed preparation.
3.2. Lignocellulosic biofeeds
All forms of the O-containing groups, i.e., alcohols, phenols, aldehydes, ketones, carboxylic acids, esters and ethers, can be present
in biofeeds of lignocellulosic origin. For more complex structures,
more than one of these groups may be part of the same molecule. In
this regard, the best known are methoxy phenols, generally termed
as guaiacol and alkyl guaiacols. The structures of compounds identied in the volatile fraction from pyrolysis of wood chips are shown
in Fig. 3 [66,67]. These structures were assigned on the basis of an
extensive GCMS evaluations. Several models of lignin, an important constituent of lignocellulosic biomass, may be found in the
literature [68,69]. Fig. 4 [69] shows the model proposed by Jongerius et al. [69] as a representation of soft wood lignin. Based on
this structure, the complexity of the products from biomass conversion shown in Fig. 3 may be better comprehended. It was indicated

20

E. Furimsky / Catalysis Today 217 (2013) 1356

Table 9
Composition of triglycerides fatty acid (wt.%) in selected vegetable oils (Cx/y ; x number of carbons; y number of double bonds).

Fatty acid
Corn oil
Linseed oil
Rapeseed
Peanut oil
Soybean oil
Sunower oil
Camelina oila
Hemp oil
Tung oil
Jatrophab

14/0

16/0

16/1

18/0

18/1

18/2

18/3

20/0

22/0

Myristic

Palmitic
12.2
7.0
1.7
11.6
11.0
6.8
7.8

Palmitoleic
0.1
4.0
63.7
0.2
4.0
0.1
3.0

Linolenic
0.9

Arachidic
0.1

Behenic

1.5
0.1
0.4

3.0

0.3
0.5

2.1
17.0

Oleic
27.5
15.0
14.3
46.5
53.2
18.6
23.1
54
14.6
47.3

Linoleic
57.0
35.0

3.1
11.3

Stearic
2.2
39.0
15.4
3.1
23.4
4.7
16.8
9
11.2
12.8

4.7

0.6

4.3
0.1
0.1
0.2

31.4
7.8
68.2
31.2
26
69.0

Ref. [14].
a
Also contains 12% C20:1 and 2.8% C22:1 [43].
b
Also contains 4.4% C24:0 [44].

above that the liquids from liquefaction of a similar biomass are


more aromatic and contain less oxygen because of an extensive
decarboxylation, deesterication and dehydration of the species
shown in Fig. 4 [69]. The extensive characterization of the biocrude
obtained by fast pyrolysis of two different lignins conducted by de
Wild et al. [70] is in line with the observations made in other studies
[6669].
Maggi and Delmon [35] compared the properties of liquids produced by fast-pyrolysis with those from slow-pyrolysis. The former
liquids had a higher content of oxygen. Carboxyl, carbonyl, methoxy
and phenolic groups accounted for most of the oxygen present. As
determined by GCMS, total of 63 oxygenates were identied in
the biofeed from fast pyrolysis of rice husk [36]. The 1 H and 13 C
NMR spectroscopy characterization of the bio-oils produced by fast
pyrolysis of biomass such as switchgrass, alfa stems, corn stover
and guayule (whole plant and latex-extracted bagasse) revealed
high complexity of aromatic compounds present [37]. Thus, the
ratio of the substituted aromatic carbons to unsubstituted aromatic carbons was greater that 2. For a longer contact pyrolysis, the

oxygen content in the primary liquids was lower because of


secondary (decarboxylation, dehydration and condensation). Generally, HPR conditions must be chosen to ensure the oxygen content
decrease below 1% [38,71,72]. The primary liquids from slowpyrolysis are less polar and more aromatic with a lesser aromatic
ring substitution. They also contain less water than the products
from fast-pyrolysis. Although the quality of liquids from slowpyrolysis seems to be better, their yield is much lower compared
with that from fast-pyrolysis.
3.3. Biofeeds of algae origin
Hydrocarbons such as n-alkadienes, trienes, triterpenoid and
tetraterpenoid, were identied in the products from the hydrothermal liquefaction of B. braunii micoalgae [51], while the biofeed
obtained from the Enteromorpha prolifera macroalgae [42] was a
complex mixture of oxygenates (e.g., ketones, aldehydes, phenols,
alkenes, fatty acids and esters), aromatics and N-heterorings. Acetic
acid was the main component of the water-soluble products. For

Fig. 3. Structures of components identied in wood pyrolysis liquids.


Ref. [66].

E. Furimsky / Catalysis Today 217 (2013) 1356

21

Fig. 4. Tentative structure of soft wood lignin.


Ref. [69].

biooils compared in Table 6 [44], the main components included


aromatic hydrocarbons, N-heterocyclics (e.g., pyrroles, indoles and
pyrrolidines), amides, amines, carboxylic acids, esters, ketones,
and straight chain hydrocarbons. A higher abundance of nitrogen heterocyclic compounds was attributed to the high content of
chlorophyll and protein in the algae biomass. Generally, pyrolysis
bio-oils had a higher content of nitrogen and aromatics than those
from liquefaction.
The oil obtained from high protein content Spirulina by
hydrothermal liquefaction (300 C, 1012 MPa, 30 min) contained O-decylhydroxylamine (11.8%), tridecene (4.7%), dimethyldioxaspiro-undecanone, (4.6%) and trimethyl-dodecanol (4.0%)
[73]. Only minor quantities of straight-chain hydrocarbons were
present, i.e., pentadecene (0.5%), heptadecene (0.9%), and docecane
(0.4%). The GC distillation showed that the Spirulina biocrude contained about 20% of 343 C fraction, 50% of VGO (343538) and 20%
residue (+538 C).
Complex mixture of products in bio oil from hydrothermal
liquefaction of brown macro-alga Laminaria saccharina, was
reported by Anastasakis and Ross [74]. Both aromatic (toluene,
ethylbenzene, and styrene) and aliphatic (1-pentadecene and
cycloalkanes) hydrocarbons were observed as well as complex oxygenates, i.e., 4-methyl-3-cyclohexene-1-carboxaldehyde,
cyclopentenone derivatives, 1-(1-cyclohexen-1-yl)-ethanone and
6,6-dimethyl-cyclooct-4-enone. These oxygenates were formed
from the carbohydrate fraction of algae, while saturated fatty acids
(e.g., tetradecanoic and n-hexadecanoic acid) were formed from the
lipids fraction.
Characterization of the bio oil from hydrothermal liquefaction of marine Nannochloropsis sp. microalga was undertaken by
Brown et al. [75]. The N-containing compounds included indole
and methyl indole. The pyrrole ring was also present in chlorophyll
suggesting that the indole formation may result from the hydrolysis

of chlorophyll in addition to two other major products such as


phytane and phytene. The major free fatty acids included myristic, palmitoleic, palmitic, oleic, and stearic acids. In comparison,
the major fatty acids obtained by hexane extraction of dry Nannochloropsis sp. included eicosapentaenoic acid, palmitoleic acid,
palmitic acid, myristic acid and arachidonic acid. They accounted
for almost 60% of the fatty acid content. The biofeed obtained from
the algae also contained cholesterol and related compounds which
arose from the lipid fraction of the algal biomass. Heptadecane, presumably formed via decarboxylation of stearic acid, was observed,
except at 500 C. The formation of polyaromatics began at 300 C
and increased with increasing temperature of hydrothermal liquefaction until they became dominant species at 500 C.
Perhaps, the most detailed characterization of the biocrude
obtained during hydrothermal liquefaction (350 C; 1 h) of several algae was conducted by Biller and Ross [76] with focus on
the effect of either 1 M Na2 CO3 and 1 M HCOOH or pure distilled water. The high protein containing cyanobacteria Spirulina
gave higher proportion of protein derived pyrroles and indoles but
also some hepta- and hexa-decane while aliphatic amides were
present under water and formic acid. The amino acid asparagine
processed in Na2 CO3 gave large amounts of phenols and some
indole compounds, however, in water and formic acid, the indoles
were converted to pyrrolidones and cyclic ketones. Bio-crude
obtained from Chlorella algae (with Na2 CO3 ) contained phenols
and piperdine derived compounds as well as alkanes such as
hepta- and hexa-decanes. Dimethyldecanamide, dodecamine and
octanoic acid were also formed in water and HCOOH. In Na2 CO3 ,
high lipid content nannochloropsis gave large amount of fatty acids
and heterocycles such as pyrroles indoles. General trends showed
that the higher the protein contents algae, the higher fraction of
pyrroles and indoles in the bio-crude. The liquefaction of soya protein and albumin in water and formic acid showed very similar

22

E. Furimsky / Catalysis Today 217 (2013) 1356

chromatographic ngerprints compared with a small difference for


the products obtained with Na2 CO3 . Under all three conditions,
phenols and indoles were the most abundant compounds. Piperdine and pyrrolidinone were also found in water and formic acid
but not in the presence of Na2 CO3.
3.4. Biofeeds from sewage sludge
A signicant variability in the composition of biofeeds derived
from sewage sludge may be anticipated. Compared with other
sources, information on the composition of biofeeds of sewage
sludge origin is rather limited. During detailed characterization of the sewage sludge derived biofeed conducted by Izhar
et al. [55c], N-heterocyclics such as pyridines, pyrroles, pyrazines,
indoles, quinoline, isoquinoline, carbazoles, harman and norharman were identied as the most abundant group of nitrogen
compounds. Less abundant were compounds formed during the
pyrolysis of amino acids, i.e., pyridinones, succinimide, pyrrolidinedione and pyrazinediones in addition to aromatic amides,
amines and nitriles. Sulfur compounds usually contained multi
sulfur atoms (e.g., dimethyltrisulde) and bi-heteroatoms such as
4-mercapto-4-methylpentan-2-one, 5-mercapto-5-methylhexan3-one, benzothiazole, etc. Different composition of biofeed derived
from pyrolysis of municipal sewage sludge was reported by Rushdi
et al. [55d]. In this case, n-alkanoic acids, n-alkanols and sterols
were the major compounds with hydrocarbon yields increasing
with increasing temperatures from 200 to 350 C. Oxalic acid in
the mixture with sewage sludge increased the concentration of
n-alkanes in liquid products.
3.5. Biofeeds from FTS processes
The liquid products from FTS are dominated by straight chain
hydrocarbons with olenic portion from high temperature FTS process being much greater than that from the low temperature FTS
process [59]. In addition, a high concentration of aliphatic oxygenates and small amount of water are always present. Sulfur and
nitrogen are not present. A much greater average molecular weight
and predominantly straight n-alkanes structure of hydrocarbons
in the products from low temperature FTS indicate that a heavier
portion of these products are in the form of solid wax.
4. Catalysts for hydroprocessing of biofeeds
Conventional catalysts consisting of Mo and/or W and Co and
Ni promoters have been used for HPR of biofeeds [7783]. The
-Al2 O3 , varying widely in textural properties (surface area, pore
volume and pore size) has been predominant. More recently, other
supports (e.g., SiO2 , TiO2 , mixed oxides, zeolites, mesoporous silica
based supports, carbons, etc.) have been tested. The unsupported
catalysts containing the same conventional metals have been tested
as well. Unconventional catalytic phases include metal carbides,
nitrides, phosphides and borides containing conventional metals
and other transition metals as well as noble metals [84,85]. They
may be used either in a supported or unsupported form. Noble
metals supported on various supports have been receiving attention as well, particularly as part of bifunctional catalysts.
4.1. Conventional HPR catalysts
The extensive information on various aspects of conventional
HPR catalysts has been periodically reviewed by several authors
[38,7796]. In the operating form, conventional HPR catalysts
contain slabs of the Mo(W)S2 crystallites. Under HPR conditions,
coordinatively unsaturated sites (CUS) and/or sulfur anion vacancies exposing metallic sites, may be created. The double and even

multiple vacancies may be formed. Many HPR reactions have been


interpreted in terms of either absence or presence of the CUS on
catalyst surface. Involvement of other types of active sites was
supported by experimental observations [81]. Generally, CUS were
much better geometrically dened than other site, which was much
less poisoned by N-bases than CUS. This was supported by the
results on HDS of DBT in the presence of pyridine [97,98]. Thus, the
HDS activity of catalyst declined by increasing amount of poison in
the feed before the curve tailed off indicating little poisoning. The
non-poisoning region was dominated by HYD reactions. Generally
accepted dual-site hypothesis was later disputed by the experimental data published by Prins [81] indicating that even for a single
heteroring compound as many as four catalytic sites may be necessary before heteroatom can be eliminated. For example, for the
HDN of quinoline, each intermediate in the overall HPR scheme
may require a different catalytic site [7981]. This would imply that
in overall HPR scheme, an adsorptiondesorptionreadsorption
and/or reorientation of the reactant intermediates would be necessary to account for a multi-site involvement. This rather complex
process may be simplied by allowing free and rapid migration of
active hydrogen on catalyst surface to ensure a gradual conversion
of reactant to hydrocarbons in consecutive reactions occurring on
and/or in a proximity of the same active site.
Understanding of the structure of active phase in HPR catalysts
was advanced using the results obtained by Mossbauer emission
spectroscopy, STM and DFT methods [83,99101]. Thus, a strong
evidence in support of the Co Mo S phase in sulded CoMo/Al2 O3
catalysts was provided. Similar structures were conrmed in the
sulded NiMo/Al2 O3 , CoW/Al2 O3 and NiW/Al2 O3 catalysts, e.g.,
Ni Mo S, Co W S and Ni W S, respectively. In this phase, an
enhanced concentration of Co and/or Ni promoters at the edge
planes of the MoS2 /WS2 crystals has been conrmed. The structure and morphology of the Co Mo S crystallites can be inuenced
by the temperature of catalyst sulding [99101]. Thus, the Type
I phase formed at lower temperatures, was still chemically bound
with the support. The occurrence of this phase was an indication of
the incomplete sulding. The sulding at higher temperatures facilitated the transformation of the Type I phase into Type II phase.
Consequently, most of the Al O Mo entities disappeared. The
addition of promoters such as Co(Ni) decreased the temperature
at which this linkages could be broken [102,103].
Signicant advancements in the understanding of the structure,
morphology, activity and selectivity of MoS2 , Co(Ni) Mo S and
Co(Ni) W S phases made by Topsoe and coworkers [104112]
claried some early opinions. These authors used a combination of
the novel experimental methods (e.g., TPD, TPS, STM, HAAD-STEM,
etc.) complemented by DFT calculations. Both an ab initio approach
and that with an incorporation of some experimental data were
used for the latter method. These studies revealed that the truncated triangle, mostly of a single layer type, was the most common
regular shape for the unpromoted MoS2 (WS2 ) clusters. The MoS2
morphology was inuenced by composition of the atmosphere. For
the graphite supported MoS2 , the shape of the clusters was inuenced by irregularities of the support surface. In this case, the MoS2
clusters were imaged as bright patches (ca. 1100 nm in diameter),
forming thin sheets with a thickness of 510 nm. The sulfur protrusions on the edges were out with respect to the S ions on the basal
plane of the MoS2 . The former exhibited a distinct double periodicity. The combination of STM method and DFT calculations revealed
the existence of an electronic edge state identied as very bright
one dimensional brim extending along the perimeter of the MoS2
nano-clusters. This state arose from perturbation of the electronic
structure near the edges relative to the interior part of the clusters. The DFT study on the HDS of thiophene under HPR conditions
also showed that the HYD pathway can take place on the Mo-edge
brim sites without requiring a vacancy [106,107] contrary to the

E. Furimsky / Catalysis Today 217 (2013) 1356

23

Fig. 6. Effect of temperature on HDN of OPA and HDS of DBT (A and B HDS of DBT
on MoS2 /Al2 O3 and sulded NiMo/Al2 O3 , respectively) at 5 MPa and 3 kPa of DBT.
Ref. [117].

Fig. 5. Top and side view of MoS2 slab (Sc , Sb and Se sulfur corner, basal and edge,
respectively; V sulfur vacancy).
Ref. [92].

DDS pathway which may involve S-vacancies on the edges [108].


Pyridine molecule interacted weakly with the MoS2 nanoclusters
[110,111]. However, when SH moieties were present, it was protonated to pyridinium ion which was stable at the Mo-edge. But,
the strongest binding occurred when the nitrogen atom of pyridinium ion was located between two sulfur atoms. The formation of
pyridinium ion depleted active surface hydrogen. Consequently,
the rate of HYD reactions was affected. Strong adsorption of pyridinium ion on brim sites slowed down the adsorption of other
reactant molecules as well.
The addition of Co and Ni changed morphology of MoS2 toward
being more hexagonal, i.e., heavily truncated triangles [112]. The
hexagonal shape of the promoted MoS2 nanoclusters suggested
the presence of two types of low-index edge terminations of the
basal plane in MoS2 , i.e., low-index (1 0 1 0) S-edges and Mo-edges
(1 0 1 0) (Fig. 5) [92]. The latter edge was inert to promoters which
could be located only on the 1 0 1 0 S-edges. These edges were energetically stabilized relative to the Mo-edge by incorporating Ni and
Co atoms. Two different metallic edge states were observed in the
NiMoS system compared with only one in Co Mo S systems.
One was similar to that in Co Mo S system, but the other was
different [113115]. The latter was responsible for differences in
catalytic activity and selectivity between Co Mo S and Ni Mo S
phases.
The brim site model can explain more consistently inhibition
and poisoning effects on HYD reactions than other models. Moreover, brim sites do not adsorb H2 S. This explains little inhibition
of HYD reactions by H2 S. They also explain the absence of steric
hindrance of alkyl substituted heterorings, e.g., during the overall HDS of 4,6-DMBT, which is dominated by the HYD pathway.
Thus, because they are very open, the adsorption of the refractory,

sterically hindered molecules can be readily facilitated. Brim sites


and their neighboring protons can interact strongly with basic Ncontaining molecules. For example, the pyridinium produced by
this interaction is much more strongly adsorbed on brim site than
pyridine alone.
It is believed that the brim sites model represents highlight
of HPR catalysis in recent years. Because of unique approach, the
authors [111115] were able to describe the most early stages and
intimate transition state of the reactions of thiophene. This has
never been accomplished before. The information on the reactions
of more complex molecules is desirable to enhance the validity of
the brim sites model, although as it is indicated in other chapters,
the observations made for large reactants may be interpreted by
involvement of brim sites.
With respect to HPR of biofeeds, H2 O may have pronounced
effect on brim sites. Thus, a possible replacement of sulfur with oxygen would change electronic character of brim sites due to different
electronegativities between sulfur and oxygen atoms. It is noted
that this issue has not yet been addressed. Yet, in the study on HDO
of phenols over unsupported MoS2 in the absence of sulfur donating agent in the feed, a continuous decline in activity was noted
[116]. However, the structural changes incurred by MoS2 were not
addressed.
The results published by Hrabar et al. [117], on the simultaneous HDS of DBT and HDN of o-propylaniline over the sulded
Mo/Al2 O3 and NiMo/Al2 O3 were interpreted in terms of the coexistence of CUS and brim sites. It was evident, that the HYD path
during HDN, giving hydrogenated hydrocarbons such as propylcyclohexene and propylcyclohexane, via o-propylcyclohexylamine
intermediate, was catalyzed only by brim sites. Little adverse effect
on this reaction was observed in the presence of the H2 S generating
DBT over MoS2 . At the same time, a dramatic enhancement in the
overall rate of this reaction was observed over NiMo/Al2 O3 (Fig. 6)
[117]. These authors suggested that higher intrinsic rate could be
related to the higher electron density at the brim sites in the presence of Ni. It may be postulated that the HDO of o-substituted
phenols may proceed in a similar way as that of the HDN of opropyl aniline. Thus, the HYD of aromatic ring occurring on brim
sites may be a dominant route proceeding via alkyl cyclohexanol to
alkyl cyclohexene and alkyl cyclohexane. Fig. 7 showing the interaction of phenol with CUS to give alkyl benzene, was adapted from
the study of Hrabar et al. [117]. This pathway is less evident relative
to HYD route.
Some experimental evidence during the preparation and testing of conventional HPR catalysts was interpreted in terms of the
coexistence of CoMoS phase with CoMoC phase [118121]. This

24

E. Furimsky / Catalysis Today 217 (2013) 1356

Fig. 7. Adsorption modes of alkyl phenol on MoS2 surface.

was later disputed on the basis of detailed spectroscopic evaluation


backed by DFT calculations [122,123].
Surface area, pore size and pore volume of HPR catalysts deserve
attention as well [8493,124]. Thus, to ensure efcient utilization
of catalyst, sufcient pore diameter, i.e., necessary to allow reactant molecules to access active surface in the interior of catalyst
particles. This is crucial for conversion of high molecular weight
species (e.g., triglycerides, lipids, proteins, etc.). Then, the overall
rate of conversion of large molecules to hydrocarbons may be determined by diffusion into catalyst particles interior. Once glycerides
are converted to free fatty acids, the diffusion effects on the subsequent conversion of the latter should be less evident. The diffusion
problem may be overcome by conducting the HPR of glycerides
containing biofeed in several stages. For example, the conversion
of glycerides in the rst stage could be maximized by employing a
large pore diameter catalyst, while a catalyst exhibiting high HDO
and decarboxylation activities would be employed in the second
stage.
Diffusion limitations should be the least evident for biofeeds
produced by pyrolysis of any biomass. On the other extreme, a special attention would have to be paid during the catalyst selection
for biofeeds obtained either from vegetable oils or algae sources
produced by non-thermal means (e.g., extraction). The biofeeds
obtained by hydrothermal liquefaction should be less depolymerized than those obtained from the same biomass by pyrolysis.
4.1.1. Modication of active phase during HPR of biofeeds
Catalyst performance during the HPR may be affected by large
amounts of water in biofeed. This may be prevented by addition of
an H2 S generating agent to biofeed. Indeed, a steady catalyst performance may be achieved but not without complications caused
by a potential contamination of the products from HPR of biofeeds
by sulfur. Rather unique case may be the biofeeds of an algae origin.
Thus, for these biofeeds, large amounts of ammonia formed during
the HDN of nitrogen containing reactants may modify the surface
structure of catalysts. The performance of HPR catalysts may also
be inuenced by a continuous coke deposition in the course of
operation.
4.1.1.1. Effect of H2 O. Contradicting information on the effect of
water on catalyst activity may be found in earlier literature [38].
Thus, water had a slight inhibiting effect on the HDO rate of

phenolic compounds and real feeds, while it had a slight promoting effect in HDN reactions. A close examination of these studies
revealed that most of the catalyst activity measurements were performed at relatively short time on stream. Also, the suldation of
catalysts may not have been complete.
Detailed study on the effect of water on structure and morphology of the unpromoted MoS2 and sulded CoMo/Al2 O3 was
conducted by Badawi et al. [125]. For the former, the exchange
of a large fraction of edge sulfur by oxygen occurred while for
CoMo/Al2 O3 , the effect of water was much less evident. This conrmed that Co atoms prevent sulfuroxygen exchange. The HRTEM
measurements revealed signicant decrease in the average length
of MoS2 slabs of the unpromoted catalyst compared with little
change for the promoted CoMo catalyst. This suggests that more
sites for the exchange of sulfur with oxygen were available on the
former catalysts.
The HDO of 2-methyl phenol was used as the model reaction to
describe the effect of water on the performance of the unsupported
MoS2 and sulded CoMo/Al2 O3 catalysts [126]. It was established
that this reaction proceeded via two pathways, i.e., the HYD pathway giving ethyl cyclohexane and the direct deoxygenation (DDO)
pathway yielding ethyl-benzene [38]. The effect of water addition on the two pathways was different. For both unpromoted
and promoted catalysts, the HYD pathway was much more sensitive to water than DDO. According to Fig. 8ac [126], HYD activity
decreased by a factor of about 3 and 1.3 for the Mo catalyst and
CoMo catalyst, respectively, while for both catalysts, little effect of
water on the DDO pathway, was observed.
Senol et al. [127] investigated the effect of water on the activity of sulded NiMo/Al2 O3 and CoMo/Al2 O3 catalysts during the
HDO of methyl heptanoate. In addition, the effect of H2 S alone
and together with water was determined. Under the same conditions, the NiMo/Al2 O3 catalyst exhibited higher activity than CoMo
catalyst. The ester conversions decreased with increasing amount
of water in the feed. Over NiMo/Al2 O3 catalyst, the conversion
increased to the same level when H2 S and water were added simultaneously compared with a higher conversion on the CoMo catalyst.
However, the conversions were highest in the presence of only
H2 S. Under all conditions, the conversions decreased with time on
stream. In parallel, sulfur content of the catalysts decreased, i.e.,
for the NiMo catalyst, it dropped from 6.9 to 6.0 wt.%, while for the
CoMo catalyst to 5.0 wt.%. At the same time, the carbon content

E. Furimsky / Catalysis Today 217 (2013) 1356

Fig. 8. Effect of partial pressure of H2 O on activity;  CoMo/Al2 O3 ,  Mo/Al2 O3 .


Ref. [126].

of the catalysts increased up to 6.3 and 7.5 wt.% on the NiMo catalyst and CoMo catalyst, respectively. On both catalysts, the total
yield of the C7 and C6 hydrocarbons decreased with the amount
of added water, while the concentrations of the oxygen-containing
intermediates increased.
4.1.1.2. Effect of sulfur. Sulfur in biofeeds counteracts the adverse
effect of H2 O on conventional sulded catalysts. However, formation of new sulfur containing compounds may not be avoided.
Sulfur in liquid products may also include dissolved H2 S and unconverted sulfur containing agent. This suggests that additional step
would be necessary to decrease the sulfur in biofuels to the levels
required by the ULSD specications. Otherwise, catalytic phases
other than those present in conventional HPR catalysts would have
to be developed.

25

Senol et al. [126] used phenol as the representative of cellulosic biofeeds and aliphatic carboxylic acid, ester and alcohol typical
of components present in the biofeeds of vegetable origin. In the
overall mechanism of the conversion of phenol over NiMo/Al2 O3
catalyst, both DDO and HYD reactions were inhibited by H2 S,
whereas over CoMo/Al2 O3 mainly DDO was inhibited. Under similar conditions, H2 S had a promoting effect on the HDO of aliphatic
oxygenates. These observations were attributed to different structures of aromatic and aliphatic oxygenates.
While using phenol and anisole over the sulded CoMo/Al2 O3
catalyst, Viljava et al. [128] observed that H2 S strongly inhibited
the hydrogenolysis to aromatics, whereas the pathway involving
HYD followed by hydrogenolysis to produce alicyclic compounds
remained similar at moderate H2 S concentrations. This was consistent with the involvement of brim sites. The demethylation of
anisole to phenol occurred even on the Al2 O3 . This reaction was
enhanced over both oxidic and sulded CoMo/Al2 O3 catalyst.
Ryymin et al. [129] reported that the conversion of esters and
alcohols over the sulded NiMo/Al2 O3 catalyst was signicantly
greater compared with that over the unsulded (reduced) catalyst.
The sulfur-containing compound such as heptanethiol was formed
during the HDO of methyl heptanoate over the sulded catalyst.
This was attributed to the presence of SH moieties formed on the
catalyst surface during sulding. It was proposed that two independent routes were involved in the formation of heptanethiol, i.e., one
involving corresponding alcohol and the other alkene, both reacting
with the SH groups. Contrary to these observations, sulfur containing compounds were not formed over the reduced NiMo/Al2 O3
catalyst.
A signicant difference between the effect of H2 S and CS2 on
catalytic activity was reported by Senol et al. [130]. Thus, for
the NiMo/Al2 O3 and CoMo/Al2 O3 catalysts, with the addition of
1000 ppm H2 S, the HDO conversion of methyl heptanoate increased
from 67 to 76% and 34 to 55%, respectively. On the other hand,
the addition of CS2 had little effect on total conversion but it
decreased the HDO conversion. For example, 3700 ppm CS2 in the
feed decreased the HDO conversion of the methyl ester from 67
to 44% and from 34 to 19% on the NiMo/Al2 O3 and CoMo/Al2 O3 ,
respectively. It is believed that the adverse effect of CS2 on HDO
conversion may be reversed by selecting an optimal H2 pressure.
Otherwise, a portion of carbon from CS2 may be left behind on surface and cause the catalyst deactivation. The study Senol et al. [130]
on the effect of H2 S and CS2 on the HDO of methyl heptanoate was
expanded to include ethyl heptanoate, heptanol and heptanoic acid
over the same [131]. In contrast to CS2 , H2 S had promoting effect on
HDO of the oxygenates by stabilizing the catalysts performance and
increasing the yield of C7 relative to C6 hydrocarbons, similarly as
observed in the above study [130]. Apparently, H2 S enhanced acidcatalyzed reactions such as hydrolysis, esterication, dehydration,
E2 elimination and SN 2 nucleophilic substitution and decreased
coke formation. Both H2 S and CS2 decreased HYD activity of the
NiMo catalyst compared with little adverse effect on the CoMo
catalyst.
4.1.1.3. Effect of ammonia. There has been no experience with the
HPR of feeds containing more than 1 wt.% of nitrogen. Yet, nitrogen
content in some biofeeds may be several times higher. Under such
conditions, signicant amount of NH3 can be produced. Then, the
interaction of NH3 with catalyst surface cannot be ruled out. The
following set of tentative reactions involving NH3 may be proposed:
Me SH + NH3 = NH4 + + Me S

(R1)

Me H + NH3 = NH4 + + MeQ

(R2)

Me SH + NH3 = NH4 SH +

(R3)

MeQ

26

E. Furimsky / Catalysis Today 217 (2013) 1356

Fig. 9. Mass balance during HDO of guaiacol ( CoMo/Al2 O3 ; CoMo/AC).


Ref. [133].

NH4 + + H2 S = NH4 SH + HS
MeQ + HS =

Me SH

(R4)
(R5)

where Q is sulfur vacancy. Reactions (R1)(R3) may be responsible for reversible catalyst deactivation due to the removal of active
hydrogen from the surface. The reaction (R4) may limit the availability of H2 S required to maintain conventional catalysts in an
active and stable form. The surface is partially reactivated via reaction (R5). Although rather speculative, the observation made by
Yang et al. [132], supporting the interaction of H2 S with nitrogen
bases, provides at least qualitative support for reaction (R4). The
sublimation point of NH4 SH (120 C) suggests that it may be the
source of deposits in downstream lines of catalytic systems. For
high nitrogen content biofeeds, combined effects of NH3 , H2 O and
H2 S on HPR may signify complexity of these effects.
4.1.1.4. Effect of coke. Coke deposition on catalyst surface, caused
by polymerization of unstable species, always occurs during HPR
[90]. Rapid deactivation of conventional catalysts during the HPR
of biofeeds obtained by pyrolysis of primary liquids from lignocellulosic sources was noted [38,133]. Fig. 9 [133] shows that for
the sulded CoMo catalyst used for the HDO of guaiacol, signicant decrease in catalyst deactivation was achieved by replacing the
-Al2 O3 support with carbon support. Thus, for the latter, the phenol/catechol ratio was about seven times greater compared with
that over the CoMo/Al2 O3 catalyst. Because the amount of Co and
Mo were similar on both supports, much greater amount of coke on
CoMo/Al2 O3 catalyst may be attributed to the acidity of -Al2 O3 .
The activity of catalysts tested during the HPR of model compounds such as free fatty acids are inuenced by their concentration
in the solvent used for the experiments [134]. For example, for
tall oil fatty acids, the conversion to hydrocarbons increased with
decreasing concentration of the free fatty acid in dodecane solvent. This is an indication of a self-inhibition and/or competitive
adsorption on active sites because the adverse effects could be
counteracted by increasing temperature.
The rapid decreases in the conversion levels of the methyl heptanoate and heptanoic acid during the initial stages on stream
observed by Senol et al. [127] were attributed to coke deposition
and the loss of sulfur from the catalyst surface. Deactivation of
the sulded NiMo/Al2 O3 catalyst was less pronounced than that
of the sulded CoMo/Al2 O3 catalyst. This was complemented by an
increased hydrogen consumption of the former.
Catalyst deactivation during the HPR of rapeseed-oil over sulded CoMo catalyst (310 C, H2 of 3.5 MPa) in xed-bed reactor was
inuenced by the purity of biofeed [135]. Among three rapeseed oil
feeds, the one containing phosphorus, alkalis and sulfur caused the

Fig. 10. Effect of rapeseed oil (RO) purity on deoxygenation ( oleic acid, 
waste RO,  partially rened RO,  primary rened RO, rened RO).
Ref. [135].

most extensive deactivation. Alkalis poisoned active sites, while


the effect of phosphorus was two-fold. Thus, when alkalis were
present, corresponding phosphates deposited on the front catalyst
bed. Without alkalis, the potential formation of phosphoric acid
that catalyzed oligomerization reactions leading to the formation
of carbonaceous deposits, was observed. These adverse effects of
alkalis and phosphorus in the biofeed were counteracted by sulfur
(added DMDS). Fig. 10 [135] obtained in this study shows catalyst
deactivation, as indicated by decreasing oxygen removal from the
biofeed, with time on stream.
4.2. Non-conventional catalysts
Non-conventional catalysts have been evaluated for HPR of
biofeeds with aim to avoid the need for catalyst presulding and
to ensure stability in the presence of large quantities of H2 O.
Noble metals supported on various supports have been attracting attention. In addition, novel catalytic phases such as metal
carbides, nitrides, phosphides and borides, were tested. For nonconventional catalysts, the effects of H2 O, H2 S, NH3 and coke have
received little attention.
4.2.1. Noble metals containing catalysts
For biofeeds containing little sulfur and nitrogen, catalysts consisting of noble metals (Pt, Pd, etc.) supported on either traditional
-Al2 O3 or novel supports, may be suitable. Noble metals supported on acidic supports have been used for HCR and HIS as part of
bifunctional catalysts, i.e., in upgrading of primary liquids from FTS

Fig. 11. Effect of temperature on IS and HCR conversion of n-tridecane over Pt/CaY
at 3.9 MPa.
Ref. [136].

E. Furimsky / Catalysis Today 217 (2013) 1356

Fig. 12. Acid site (OH), metallic site (Mo) and dual site model for MoOx Ny .
Ref. [138].

[5962]. Potential of bifunctional catalysts in upgrading biofeeds


from vegetable oil sources may be anticipated. In this case, they
facilitate HIS and HCR activities. The extent of HIS relative to HCR
depends on temperature in relation to the acid strength of the catalyst and the overall metal to acid site balance (Fig. 11) [136].
The evidence for catalyst deactivation was presented in the
study on coHPR of the blends of petroleum fractions and vegetable
oil biofeeds (lauric acid, Pd/C catalyst at 255 300 C) published by
Mki-Arvela et al. [137]. In this case, deactivation increased with
increasing amount of biofeeds in the mixture. The catalyst stability
could be enhanced by using lower boiling solvents, i.e., mesitylene
and decane compared with dodecane. The poisoning by the product gases such as CO and CO2 as well as coke deposition were the
main causes of catalyst deactivation. The catalyst reactivation with
hydrogen did not restore its activity to the original level.
4.2.2. Metal carbides, nitrides and phosphides
Rather peculiar trends exhibited by intrinsic activity of metal
carbides and nitrides involve the HPR activity increase with
increasing particle size and/or decreasing surface area [85]. This
contradicts generally observed trends involving conventional HPR
catalysts [83,8688]. This may be attributed to structural features
of the crystals of metal carbides and nitrides allowing diffusion of
hydrogen to sub-surfaces, followed by its activation [85]. The active
hydrogen may then migrate to the surface to participate during
HPR. This speculative interpretation needs to be experimentally
conrmed.
The information regarding the nature of active sites on metal
carbides and nitrides is rather limited. Fig. 12 shows the model
of Mo oxinitride proposed by Miga et al. [138]. Three types of
sites include acidic, metallic and dual sites. It is evident that nitrogen, carbon and oxygen atoms can be inserted into the octahedral
sites located either in the center or edges of the face-centered
cubic lattice of Mo(W) atoms. This would facilitate the formation
oxycarbonitride-like structure. This suggests that if metal nitrides
and carbides are used for HPR of biofeeds, gradual change to corresponding oxynitrides and oxycarbides caused by the presence of
H2 O, may not be avoided. However, this may be benecial, because
of a higher HPR activity observed over the latter catalysts compared with metal carbides and nitrides [85]. This is supported by the
results on performance of the bimetallic NiMo carbide supported
on SiO2 during the HDO ethyl benzoate, acetone and acetaldehyde
[139]. First of all, NiMo carbide was much more active than the sulded CoMo/Al2 O3 catalyst, but the former catalyst was modied by
incorporation of oxygen.

27

Recent attempts made by Tominaga and Nagai [140] focused on


potential active sites on Mo nitrides. In this case, DFT calculations
were used to identify the most probable route for the HDN of carbazole (CAR) on Mo2 N. The most stable conguration was attained
when the nitrogen atom of CAR was on the top position of the Mo
atom with two benzene rings along the x-axis in a boat conguration. The direct C N hydrogenolysis of CAR on the Mo2 N(1 1 0) slab,
leading to the formation of bi-phenyl was not favorable. The proposed mechanism of the HDN of CAR considered two active sites.
The successive HYD proceeded on the Mo atom and one C N bond
on the same Mo atom, while the nitrogen removal of the hydrogenated CAR and the dissociative hydrogen adsorption occurred on
the bridge position of the Mo atoms. This DFT calculation estimated
the HYD of CAR to tetrahydroCAR and octahydroCAR was required
for the C N bond scissions. Cyclohexylbenzene was formed from
the former intermediate, while cyclohexylcyclohexene and bicyclohexyl were formed from the latter. The activation energy of the
HYD step was higher than that of C N bond cleavage. In the absence
of any information on the HDO of model compounds, only a speculation on the involvement of active sites may be made on the basis
on the above observations. For example, for phenols, both HYD and
DDO routes may be predicted, while for O-heterorings, the HYD
route would be more predominant.
The information on the nature of active sites on metal phosphide catalysts is rather limited in spite of growing interest in these
phases for HPR applications. The resent work published by Zhao
et al. [141] on the HDO of guaiacol over the Ni2 P, Co2 P, Fe2 P, WP and
MoP catalysts indicated a signicant effect of the type of metal on
product distribution. For example, the absence of catechol among
the products obtained at lower contact time over Ni2 P compared
with the other metal phosphides suggested a different involvement
of metals either during hydrogen activation or the interaction of
reactants with catalyst surface. This is supported by difference in
activation energies estimated for these phosphides between 200
and 300 C, i.e., from 23 to 63 kJ/mol. An additional information is
needed to clarify these issues. However, the turnover frequency
determined by titration of active sites using the chemisorption of
CO followed the order: Ni2 P >Co2 P >Fe2 P WP MoP. This order
coincided with the activity order obtained during the HDO of guaiacol [141].
4.3. Role of support
During HPR of biofeeds, supports can play signicant role
as shown in Fig. 9 [38,133] comparing stability of the sulded
CoMo/Al2 O3 catalyst with CoMo/AC catalysts both containing the
same amount of metals. These results were obtained during the HPR
of biofeeds from pyrolysis of the lignocellulosic biomass. A rapid
deactivation of the CoMo/Al2 O3 compared with a little deactivation
over CoMo/AC catalyst was rather evident. This may be attributed
to the presence of acidic sites on -Al2 O3 surface compared with
the neutral surface on the carbon. Also, the textural properties (e.g.,
porosity and surface area) of carbon were more favorable for an
efcient distribution of active metals. However, carbon alone was
observed to exhibit HYD activity [142] and ability to store hydrogen [143]. In this regard, several active sites on the surface of carbon
were identied [144148]. The stability of supports in H2 O needs
also to be addressed. A gradual conversion of -Al2 O3 to boehmite
has been frequently observed. Under the same conditions, carbon
supports are hydrophobic.
The coke deposition on -Al2 O3 supported catalysts may be
offset by increasing supply of active hydrogen. This may be
achieved by employing active metals possessing a high HYD activity. Indeed, noble metals (Pt, Pd, etc.) containing catalysts supported
on -Al2 O3 have been successfully tested for the HDO of guaiacols, biphenols and anisole [149,150a]. In this case, a reasonable

28

E. Furimsky / Catalysis Today 217 (2013) 1356

stability of the catalysts was observed. Moreover, such catalysts do


not require presulding.
The above comparison of carbon with -Al2 O3 supports is more
relevant to the HPR of biofeeds of a lignocellulosic origin than
those derived from vegetable oil sources. For the latter biofeeds, the
presence of long chain parafnic structure, which are part of free
fatty acids requires attention. Obviously, high conversions of such
reactants to hydrocarbons can be achieved by employing either
carbon or -Al2 O3 supported catalysts [15,16]. However, desirable
cold ow properties of the hydrocarbon products (to be used as
diesel fuels) may not be attained. This problem may be overcome
by employing more acidic supports (e.g., SiO2 Al2 O3 , zeolites, etc.)
such as used in bifunctional catalysts [59].
A high nitrogen content in biofeeds from algae biomass suggests
that the selection of support for catalysts used in such applications
may be of a crucial importance because of high yields of ammonia produced. For such applications, neutral support such as carbon
may be the most suitable. Otherwise, a strong interaction of ammonia with acidic sites may diminish the role of support as the source
of proton required for HIS reactions.
5. Reactions occurring during HPR of biofeeds
A signicant diversity in composition among biofeeds from different sources was indicated in Sections 2 and 3. In HPR studies,
primary focus has always been on the components which have the
most detrimental effects on the properties of biofuels. For biofeeds
derived from vegetable oils and lignocellulosic sources, conversion
of the oxygen-containing compounds to hydrocarbons and removal
of aromatics may be the main objective. However, the presence
of N-containing compounds makes the composition of the algae
derived biofeeds much more complex compared with the other
biofeeds. Thus, as much as 10 wt.% of nitrogen may be present.
This suggests that more than half of all molecules comprising such
biofeeds contain a N-containing group.
5.1. Mechanism of HPR reactions
The understanding of mechanistic aspects of the HPR of biofeeds
benets from extensive information established during the HPR
of petroleum and other non-petroleum feeds. Because of a high
oxygen content in biofeeds, primary focus is on HDO reactions.
Historically, little attention has been paid to the HDO of petroleum
feeds because the content of oxygen rarely exceeded 0.5 wt.% [145].
Phenols and furanic heterocyclic structures have been the dominant O-containing compounds in conventional petroleum feeds
[38]. The HDO reactivity of furanic rings decreases in the following
order: furan > benzofuran (BF) > di-benzofuran (DBF). However, an
overlap in the reactivity may be evident in the case of alkyl substituted O-heterorings. General trends indicate that furan, BF and DBF
are less reactive than corresponding thiophenic analogs, although
this distinction becomes less clear when different temperatures,
H2 pressure and alkyl substitution are taken into consideration.
In the case of biofeeds, the presence of furanic structures have
been rarely reported, although their presence cannot be completely
ruled out. The presence of naphthenic acids in some petroleum
feeds attracted attention because of their corrosive nature. Naphthenic acids may also be present in some biofeeds.
In this sub-chapter, focus is on the O-containing compounds,
aromatic structures and N-containing compounds the presence of
which was clearly conrmed by the extensive biofeeds characterization. The discussions of mechanistic aspects are organized
according to single components and the corresponding biofeed in
which these components are most abundant. It should be noted
that studies on mechanism of HPR reactions involving model

Fig. 13. Simplied mechanism of conversion of guaiacol to hydrocarbons.

compounds have been an essential source of the information for


elucidating HPR mechanism.
5.1.1. Reactants of ligno-cellulosic origin
Various forms of o-methoxyphenols referred to as guaiacols,
anisoles and bi-phenols, as the most abundant species, have
been identied in biofeeds of lignocellulosic origin. Less abundant species include ethers, carboxylic acids and esters. The origin
of these compounds may be deduced from the model structure
of lignin (Fig. 4) [69], an important constituent lignocellulosic
biomass. Indeed, the extensive evaluations of the bio-feed obtained
by pyrolysis of lignocellulosic biomass conducted be Maggi and
Delmon [66,67] are in line with these expectations as shown in
Fig. 3.
5.1.1.1. Guaiacols. Guaiacols are the compounds of primary interest because of their abundance, resistance to the overall HDO and a
high coke forming propensity. Under typical HPR conditions, other
structures (e.g., acetophenone, aldehydes, alcohols, etc.) are much
more reactive and are converted to hydrocarbons while the overall
HDO of the stable species is still only in the early stages [38].
The simplied mechanism of the HDO of guaiacol is shown in
Fig. 13 [146,147]. Usually, the rst step involves hydrogenolysis of
the CAL O bond. At the same time, the hydrogenolysis of the CAR O
bond was barely noticeable. The overall HDO then proceeded via
stepwise HDO of bi-phenol to phenol and nally to hydrocarbons.
Therefore, some of the events occurring during the HDO of guaiacol are common with those observed during the HDO of phenols.
However, rather low stability of the guaiacol reactant and bi-phenol
intermediates compared with phenol should be noted. The former
compounds are the main cause of excessive coke formation unless
the HPR conditions are optimized.
Similarly as over conventional HPR catalysts, the HDO of guaiacol over Pt/Al2 O3 at 573 K (near atmospheric H2 ) took place
with a high selectivity for CAL O bond hydrogenolysis [149].
The most abundant products included phenol, catechol and 3methylcatechol. Less evident were methyl group transfer reactions
leading to the formation of alkylated phenols, guiacaols and
benzene. One of the pathways for removal of phenol involved
Table 10
Effect of temperature (K) on product distribution during conversion of guaiacol over
Pt/Al2 O3 .
Product

523

573

623

Benzene
Toluene
Anisole
Cyclohexanone
Phenol
O-cresol
Veratrole
3-Methylguaiacol
6-Methylguaiacol
Catechol
3-Methylcatechol

0.004
0.003
0.02
0.1
0.4
0.01
0.05
0.005
0.04
0.3
0.03

0.002
0.0004
0.009
0.03
0.3
0.03
0.02
0.01
0.03
0.4
0.1

0.002
0.0005
0.008
0.01
0.3
0.03
0.02
0.02
0.03
0.3
0.3

Ref. [150a].

E. Furimsky / Catalysis Today 217 (2013) 1356

29

Fig. 14. Detailed mechanism of conversion of guaiacol to intermediates and hydrocarbons.


Ref. [150a].

cyclohexanone intermediate. Perhaps, the most detailed mechanism for the conversion of guaiacol shown in Fig. 14 [150a], was
developed on the basis of an extensive characterization of products. Rather mild conditions employed (e.g., near atm. pressure
of 30% H2 /70% N2 at 573 K) ensured the presence of numerous
primary products and intermediates, which otherwise would be
undetectable under more severe conditions. The most abundant
products and intermediates detected under these conditions are
listed in Table 10 [150a]. As one would expect, the alkyl transformation reactions were evident over Pt/H zeolite [150b] as well
as over zeolite alone [149]. However, the latter was inactive for
oxygen removal.
A signicant difference in the mechanism of HDO of guaiacol over Rh/ZrO2 catalyst compared with that over sulded
CoMo/Al2 O3 and NiMo/Al2 O3 catalysts was reported by Lin et al.
[151]. For the former, the rst step involved the HYD of aromatic
ring in guaiacol, followed by demethoxylation and dehydroxylation, whereas for the conventional catalysts, mechanism involved
usual steps shown in Fig. 13 [146,147]. For anisole, limited HDO
was observed over sulded ternary CoMoW catalysts supported on
SBA-15 and SBA-16 (310 C; 3 MPa) [152a]. These catalysts exhibited a bifunctional activity as indicated by demethylation yielding
phenol, followed by isomerization to o-cresol and o-xylenol, as the
main reactions.
An allylated guaiacol such as eugenol (4-allyl-2methoxyphenol) is among usual compounds in pyrolysis
biofeeds. Under mild conditions (e.g., Pt/Al2 O3 , 573 K, near
atm. H2 , continuous system) used by Nimmanwudipong et al.
[152b], the most abundant product included 4-propylguaiacol,
while among the less abundant products, cis-isoeugenol,
trans-isoeugenol, 4-prop-2-enylphenol (p-allylphenol) and
4-propylphenol, were noted. Because of the mild conditions
applied, only traces of phenol, 4-propylbenzene, cresols, guaiacol,
catechol, methylguaiacols, 3-methylcatechol, p-vinylguaiacol,
1,2-dimethoxy-4-n-propylbenzene
(4-propylveratrole),
and
4-propylcatechol, were formed.

5.1.1.2. Anisole. In spite of the simple structure, rather complex


mechanism of the HDO of anisole, was observed [153]. Phenol as
the primary product could be formed via several pathways, each
with a different co-product such as methane, 2-methylanisole and
4-methylanisole. 2-Methylphenol and 4-methylphenol could also
be formed via intramolecular transalkylation of anisole or intermolecular transalkylation involving phenol and anisole. The multi
methyl substitution of the aromatic ring to produce dimethylphenols, trimethylphenols, and so on may involve additional pathways.
The HDO of anisole produces methanol and benzene. The formation of cyclohexanone, most likely produced via HYD of phenol
was also observed. A participation of proton in such mechanism is
highly probable. This would imply that the extent of transalkylation
reactions should increase with increasing acidity of catalysts.
Among several pathways, the mechanism of the conversion of
guaiacol in Fig. 14 [150a] involves one route leading directly to
the formation of the anisole intermediate which is subsequently
converted to the products discussed above. Similar detailed mechanisms were also developed by Runnebaum et al. [153] for the overall
conversion of anisole and Nimmanwudipong et al. [154] for conversion of cyclohexanone. The latter is found in products of a partially
converted guiacol and anisole as an intermediate during the overall
conversion.
5.1.1.3. Lignin. Horcek et al. [155] and Kleinert et al. [156,157]
developed mechanism of the HDO conversion of lignin model
molecule under typical HPR conditions. Fig. 15 [155] indicates
the formation of gaseous products, hydrocarbon liquids and
oxygenates. In addition, coke formation was evident as well. Hydrocarbon products were dominated by C6 C9 aromatics and C5 C8
i-parafns. Increasing H2 pressure resulted in the increase of the
yield of hydrocarbons relative to that of oxygenates. At the same
time, coke yield was decreased. The lignin structure in Fig. 15 [155]
shows the presence of the guaiacol precursors, similarly as another
model of lignin shown in Fig. 4 [69]. Therefore, during lignin pyrolysis, the products enriched by guaiacols should be anticipated in

30

E. Furimsky / Catalysis Today 217 (2013) 1356

Fig. 15. Mechanism of conversion of lignin under typical HPR conditions.


Ref. [155].

accordance with numerous reports in the scientic literature [38].


Guaiacols, anisole and biphenols are not included among the products in the mechanism shown in Fig. 15 [155]. Thus, under a high H2
pressure employed by Horcek et al. [155] these reactants, formed
initially in primary reactions, were converted to hydrocarbons in
secondary reactions. However, the study of Jongerius et al. [69] was
conducted under less severe conditions. In this case, numerous Ocontaining intermediates were identied. Consequently, perhaps
the most comprehensive mechanism of the HDO of reactants found
in the lignocellulosic biofeed, could be proposed.
5.1.1.4. Phenols. Phenols are always important components of the
biofeeds of lignocellulosic origin (Fig. 3) [66,67]. Phenol intermediates determine the rate of the overall HDO of guaiacols and anisole
(Fig. 14) [150a]. It was reported that it was easier to hydrogenate
phenol to cyclohexanol than benzene to cyclohexane [158]. This
was attributed to the ring electrons delocalization caused by the
CAR OH moiety of phenol [38]. In real feeds, alkyl phenols are far
predominant form of phenols with methyl being the main alkyl
substituent, particularly in the biofeeds which were subjected to a
partial HPR. Several studies concluded that the alkyl substitution
in ortho-position to OH group had the most adverse effect on the
overall HDO [38,159,160]. This is the conrmation of steric effects
during the transfer of active surface hydrogen to reactant. On the
other hand, thermodynamic considerations suggested that for alkyl
free phenol, the conversion to benzene and H2 O should proceed
readily [38]. Otherwise, kinetic factors (e.g., alkyl substitution, orientation of phenols at catalyst surface, active hydrogen availability,
etc.) inuence HDO mechanism. Also, an electron donating effect of
alkyls in m-positions may counteract the effect of CAR OH moiety

and as such have a stabilizing effect on aromatic ring. Therefore, in


spite of a relatively simple structure, the overall mechanism of the
HDO of phenols may be rather complex.
The HDO of alkyl phenols involves two main routes, presumably
initiated on different catalytic sites, i.e., a direct HDO (DDO) leading to the formation of alkyl benzenes and the intermediate HYD
of aromatic ring yielding either alkyl cyclohexanes or alkyl cyclohexenes [38]. The former route may dominate overall HDO under
active hydrogen limited availability. Under moderate H2 pressure, the DDO route dominates overall HDO over the catalyst with
a low HYD activity, whereas the HYD route over the catalysts with a
high HYD activity (e.g., noble metals suldes). On the conventional
Co(Ni)Mo(W)/Al2 O3 catalysts, the direct route requires a coordination (presumably with CUS) which can facilitate hydrogen transfer
to CAR OH, as it is shown in Fig. 7. Either at or tilted interaction of
the ring with catalyst surface (presumably brim sites) may be more
favorable for the HYD route.
Alkyl cyclopentanes sometimes observed during the HDO of
phenols, arise from aromatic ring isomerization facilitated by a
proton via oxonium ion and/or carbocation formed by H2 O elimination of the former. Carbocation may be also formed readily
by the addition of proton to a double bond in cyclohexenes at
least partially formed via dehydration of cyclohexanols [161]. The
involvement of proton during the overall HDO of phenol was considered in the mechanism proposed by Massoth et al. [162] shown
in Fig. 16. Other conclusions on mechanism of HDO of phenols
may be drawn from kinetic data discussed latter. According to
Romero et al. [163], the HDO of 2-ethylphenol (340 C and 7 MPa
of total pressure, xed-bed reactor) over Mo/Al2 O3 , CoMo/Al2 O3
and (NiMo/Al2 O3 catalyst) in the presence of DMDS involved three

E. Furimsky / Catalysis Today 217 (2013) 1356

31

Scheme 2.

Scheme 3.

Fig. 16. Mechanism of transformation of phenol.


Ref. [162].

O
\\
CH3(CH2)x - C
\

CH3(CH2)x-1CH3 + CO2
+

nH2

O-H

CH3(CH2)xCH3 + 2 H2O
Scheme 1.

pathways. The rst was the HYD of aromatic ring followed by dehydration giving 1-ethylcyclohexene and 3-ethylcyclohexene, which
were subsequently hydrogenated to ethylcyclohexane. The second
route involved direct deoxygenation (DDO) leading to ethylbenzene. Participation of surface acidity was indicated by the formation
of isomers of 2-ethyl phenols and diethylphenols via third route.
The Ni and Co promoted catalysts were much more active than the
Mo/Al2 O3 catalyst. While Ni promoted HYD pathway, Co promoted
both the HYD and DDO pathways suggesting that the DDO/HYD
selectivity was catalyst type dependent.
The oxidic supports other than -Al2 O3 may inuence the mechanism of HDO of phenol. For example, on SiO2 , phenol interacted
via hydrogen-bonding, whereas on -Al2 O3 phenol dissociated on
acidbase pairs leading to the formation of phenolates species
[164]. On the sulded CoMo/Al2 O3 catalyst, part of the phenol dissociated on the support without strongly interacting with the active
sulde phase. However, a decrease in the accessibility to sulde
sites may be observed if phenolate species is in the close vicinity
of sulde slabs. Consequently, a deactivation of catalyst may be
observed. Entirely different form of interaction may be present on
carbon supports. In this case, a C O interaction may play a key role
during the adsorption. Therefore, the type of support may be an
important factor inuencing the HDO mechanism of phenols.
5.1.1.5. Carboxylic acids and esters. Carboxylic acids are present in
almost all high oxygen containing feeds including the synthetic
liquids produced via FTS, where they are important constituents
of the oxygenates fraction [59]. Carboxylic acids may also be the
most abundant O-containing compounds in the liquids derived
from some oil shales. They are also important intermediates
during the HDO of esters found in bio-feeds. In the form of naphthenic acids, carboxylic acids are present in some petroleum feeds
[145,165,166]. For biofeeds of cellulosic origin, the carbon number of carboxylic acids is much smaller compared with that of acids
present in the vegetable oil and algae biofeed, i.e., C1 to C8 and C18 ,
respectively. A similar comparison also applies to the corresponding esters.

Fig. 3 [66,67] indicates the presence of structures in which carboxylic groups and esteric group are attached to aromatic ring. This
would suggest that corresponding naphthenic acids could play certain role as intermediates during the overall conversion of such
acids to hydrocarbons under. During HPR of biofeeds, naphthenic
acids have been rarely included in the model compounds studies.
The overall conversion of carboxylic acids to hydrocarbons over
conventional HPR catalysts may involve at least two routes occurring in parallel, i.e., one involving decarboxylation and the other the
HYD of carboxylic group to methyl group according to the Scheme 1,
i.e.,
The latter route requires the presence of at least six atoms
of active surface hydrogen in the proximity of carboxylic group
compared with one hydrogen atom consumed during decarboxylation. Therefore, the availability of active surface hydrogen may
be crucial for the former route. This may also require a special
orientation of reactant at the surface of catalyst. Nevertheless,
the study of Afonso et al. [167] conrmed that with the increasing H2 pressure, i.e., increasing active hydrogen availability, the
HYD route was gaining relative to decarboxylation and it was
predominant route at 12.5 MPa and 673 K. However, at 553 K the
HDO of decanoic acid and ethyldecanoate gave nonane and decane
with the selectivity ratio nonane/decane of 1.5 and 1.1, respectively. At 673 K and in the absence of catalyst, almost 35% of
the 1,2-(4,4 -dicarboxyphenyl)ethane in tetraline was converted
via decarboxylation. Therefore, non-catalytic decarboxylation may
occur in parallel with catalytic routes with relative contribution
depending on the experimental conditions such as H2 pressure and
temperature, as well as the type of catalyst. Because of thermal
effects involved, the contribution of decarboxylation to the overall conversion of carboxylic acids to hydrocarbons should increase
with increasing temperature [168]. Simultaneously, the formation
of CO was observed conrming the coexistence of decarboxylation
and decarbonylation reactions. The following tentative mechanism
is proposed to account for the decarbonylation reaction (Scheme 2).
The carbon balance of this transformation is only shown. Obviously,
these changes occurred under H2 pressure. Then, the products of
decarbonylation reaction should include a hydrocarbon containing
one carbon less than the parent carboxylic acid in addition to H2 O.
Aldehyde intermediate may also be present in an enolized form as
part of the enolketol equilibrium shown in Scheme 3. Therefore,
this equilibrium may be part of the overall decarbonylation process.
The HDO of esters proceeds either via decarboxylation or HYD
of carboxylic group to methyl group, after the hydrogenolysis of
ester to carboxylic acids and alcohol. Laurent and Delmon [147]
used diethyl sebacate as model compound to study the HDO of
esters. Under HPR conditions, the de-esterication (DES), i.e., conversion of the ester group to carboxylic group and an alcohol may
proceed according to tentative mechanism shown in Scheme 4. But

32

E. Furimsky / Catalysis Today 217 (2013) 1356

O
O
O
O
||
||
||
||
H2
H5C2-O-C-(CH2)8-C-O-C2H5
H5C2-O-C(CH2)8-C-OH + C2H5OH
Scheme 4.

this involves only one ester group. A very special orientation of diester molecule would be necessary to achieve a simultaneous DES
of both groups. Otherwise, a complete DES could only be achieved
in two consecutive steps. De-carboxylation of the deesteried
di-dicarboxylic acid would proceed according to the mechanism
proposed above for carboxylic acids. According to this mechanism,
nal hydrocarbons could include C8 H18 , C9 H20 and C10 H22 providing that a sufcient amount of active hydrogen was available.
This assumes the involvement of decarboxylation (decarbonylation), decarboxylation (decarbonylation)/HYD and a complete HYD
routes, respectively. The consumption of hydrogen would increase
in the same order and at least twelve atoms of active hydrogen
would be necessary to obtain C10 H22 . Eight atoms of active hydrogen would be wasted as part of H2 O rather than being incorporated
in the hydrocarbon. Alcohol, i.e., C2 H5 OH, released during deesterifecation step would be converted to C2 H6 and C2 H4 . In this case,
C2 H4 could be formed by dehydration of alcohol and subsequently
converted to C2 H6 . The -Al2 O3 alone was active for DES of esters
suggesting that support of conventional catalysts may play a more
prominent role during the HDO of esters and carboxylic acids
compared with the other HPR reactants [147]. Moreover, thermal
(non-catalytic) conversion of esters in a hydrocarbon medium was
also reported [11].
The study of de Brimont et al. [169] claried several issues
regarding the overall transformations of ethylheptanoate over
unsupported suldes such as MoS2 , Ni3 S2 and NiMoS2 . The experiments were conducted in the continuous xed-bed reactor under
HPR conditions. Over MoS2 , the reaction was dominated by HDO
yielding mainly heptane and heptenes. However, decarboxylation
(DCO) pathways, yielding hexane and hexenes was the dominant
route over Ni3 S2 . Both HDO and DCO were competitive reactions over NiMoS2 phase yielding C6 and C7 hydrocarbons. For the
NiMoS2 , the occurrence of both pathways was attributed to the
presence of Ni3 S2 in the NiMoS2 phase. This is conrmed by the
results in Fig. 17 [169] showing an increase in the DCO contribution
to the overall conversion with the increasing Ni/Mo ratio. Under
the same conditions, additional experiments were conducted using
the aldehyde intermediate heptanal. For this reactant, HDO to
hydrocarbons was the preferential route over all catalytic phases

investigated. The rate of this route was enhanced by the addition


of Ni to MoS2 . At the same time, little promoting effect of Ni on
the reaction of ethylheptanoate, was observed. Apparently, in the
presence of carboxylic acid, the HDO pathway was inhibited while
the DCO pathway enhanced.
The mechanism in Fig. 18 was adapted from the studies of
Ryymin et al. [170] and Senol et al. [171] on the product distribution
during the HDO of methyl heptanoate (69 MPa, 200 and 250 C,
00.4 wt.% DMDS, NiMo/Al2 O3 ). The reaction routes in Fig. 18
include only a gradual reductive conversion of carboxylic group.
Obviously, decarbonylation and thermal DCO occurred in parallel as well. The presence of several intermediates produced either
by gradual reduction to C7 H16 or some intramolecular reaction of
intermediates, should be noted. It was evident that the hydrolysis of ester group, followed by the reduction of carboxylic group to
aldehyde intermediates was part of the overall HDO mechanism.
However, Donnis et al. [172] reported that some observations made
during the HDO of methyl laurate did not support a simple stepwise
HYD of the aldehyde. Apparently, the enolization of the aldehyde
intermediate prior to HYD offered a more plausible explanation.
Thus, rather low reactivity of a ketone lacking -hydrogen (cannot be directly enolized) compared with a high reactivity of ketone
with -hydrogen, was noted. The keto-enol equilibrium is part of
the mechanism shown in Fig. 18. The effects of H2 S and CS2 on the
HDO of methyl heptanoate, ethyl heptanoate, heptanol and heptanoic acid over the sulded NiMo/Al2 O3 and CoMo/Al2 O3 catalysts
were addressed earlier [130,131]. The yield of alkenes (hexenes
and heptenes) increased and that of alkanes (hexane and heptane)
decreased with increasing content of DMDS in the feed.
The sulded NiMo/Al2 O3 and CoMo/Al2 O3 catalysts were used
to study reactions of methyl heptanoate, heptanol and heptanoic
acid in ow and batch reactors [127]. Methyl heptanoate produced
saturated and unsaturated C7 and C6 hydrocarbons via oxygencontaining and sulfur-containing intermediates. The latter resulted
from desulfurization of catalysts. Heptanol and heptanoic acid were
the primary products of the methyl ester. Heptanol yielded C7
hydrocarbons, whereas heptanoic acid C7 and C6 hydrocarbons.
Heptane and heptenes were the main products over the NiMo and
CoMo catalysts, respectively. The NiMo catalyst exhibited better
activity for the overall conversion of aliphatic oxygen-containing
compounds than the CoMo catalyst. This was complemented by the
higher hydrogen consumption of the former catalyst. Fig. 18 shows

Fig. 17. Effect of Ni/Mo ratio on HDO and DCO selectivity.


Ref. [169].

Fig. 18. Mechanism of HDO of methyl heptanoate.

E. Furimsky / Catalysis Today 217 (2013) 1356

33

database of thermodynamic data available in the literature is rather


limited. To overcome this problem, Smejkal et al. [178] used the
Jobacks contribution method to generate necessary data with
aim to propose mechanism of the overall conversion of glyceride
such as tristearate. The reactions (R6)(R9) below were proposed
to account for catalytic transformation, i.e., hydrodecarboxylation
(1), HDO (2), CO formation (3) and reversed steam reforming of
methane (4), i.e.,
(CH2 )2 CH[(CH2 )16 ]3 (COO)3 (CH3 )3 + 3H2 = 3C17 H36 + C3 H8
+ 3CO2

Fig. 19. Hydrogenation equilibrium of triglyceride.

that during gradual transformation of carboxylic acids and esters


to hydrocarbons, several intermediates and hydrocarbon products
always coexist. Then, intramolecular reactions involving some of
these species may lead to the formation of new compounds. For
example, a re-esterication of the alcohol with parent carboxylic
acid shown in Fig. 18, was indeed observed [170]. Other potential
reactions may include the formation of ethers from alcohols.
5.1.1.6. Other oxygenates. Other oxygen containing compounds
may also be present in biofeeds or be formed as intermediates
during the overall HDO. For example, the occurrence of alcohols,
ketones, aldehydes and anhydrides as intermediates may be anticipated. These reactants have been receiving much less attention
presumably because of their high reactivity. The formation of ethers
(alkylalkyl, alkylaryl and arylaryl) and furanic rings during the
production of biofeeds from lignocellulosic biomass should also be
expected. Database on HPR mechanism of these reactants has been
established as part of the studies involving the feeds of petroleum
origin, CDL, FTS liquids and oil shale liquids [38,173,174].
5.1.1.7. Aromatic structures. A large portion of oxygen in the
biofeeds from lignocellulosic biomass is associated with aromatic
ring suggesting that aromatics are important intermediates during the overall HPR. For example, the presence of propylphenyl
entities in lignin portion of lignocellulosic biomass has been generally conrmed. An extensive database on the HPR mechanism of
aromatic structures has been established by Klein et al. [175] and
Aubert et al. [176].
5.1.2. Reactants in vegetable biofeeds
Glycerides are the main components of vegetable oils and fats
where they are usually present in the form of triglycerides. Every
triglyceride contains three fatty acid chains esteried with a single glycerol molecule. As it is shown in Table 9 [14], the fatty
acids chains containing 18 carbons, are far predominant. The fatty
acid chains may occur as saturated, mono-unsaturated and polyunsaturated. Small amounts of C16 fatty acids (Table 9) may also be
present.
5.1.2.1. Mechanism of glycerides conversion. Under HPR conditions,
HYD equilibrium may be established between the glycerides
comprising unsaturated fatty acids chains and those with corresponding saturated chains [177]. The existence of the HYD
equilibrium is shown in Fig. 19 [177]. Such equilibrium may be
established rapidly. Then, the ratio of free fatty acids with unsaturated chains to that with saturated chains should increase with
increasing temperature and decreasing H2 pressure. Otherwise,
HDO and decarboxylation reactions determine the overall rate of
the triglycerides conversion to hydrocarbons.
Thermodynamic equilibrium based on the Gibbs energy minimization is a useful tool for developing the mechanism of HPR
reactions without performing experiments. Unfortunately, the

(1)

(CH2 )2 CH[(CH2 )16 ]3 (COO)3 (CH3 )3 + 12H2 = 3C18 H38 + C3 H8


+ 6H2 O

(2)

CO2 + H2 = CO + H2 O

(3)

CO2 + 4H2 = CH4 + 2H2 O

(4)

The predictions made by the thermodynamic model were compared with the experimental observations assuming a total HYD
of tristearate to hydrocarbons. The predictions suggested that C18
hydrocarbons were the main products and that their concentration
was inuenced by temperature and by H2 pressure. In addition
the results suggested that the reaction was limited by hydrogen
transfer. The predictions were in a reasonable agreement with
experimental data [177].
The mechanism of HDO of triglyceride proposed by Huber
et al. [168] was based on the product distribution over the sulded NiMo/Al2 O3 catalyst between 300 and 450 C, 5 MPa and
LHSV of 5 h1 . This mechanism incorporates the basic features
of that shown in Fig. 18 [170,171] except that the overall conversion of triglyceride was gradual via di- and mono-glycerides
before decarboxylation, decarbonylation, dehydration and HYD
reactions became evident. Straight chain hydrocarbons, as primary
products may continue to react via HIS and HCR reactions. The
extent of these reactions depends on the structure of catalyst,
e.g., they would be favored over bi-functional catalysts. Madsen
et al. [179] observed that the conversion of tripalmitin glyceride to
hydrocarbons proceeded via free fatty acids in agreement with the
mechanism in Fig. 18. The presence of H2 S (CoMo/Al2 O3 ; 310 C;
3.5 MPa) inuenced the overall mechanism by increasing decarboxylation pathway relative to HDO pathway [142]. At 2 MPa of H2 and
over Pt/Al2 O3 catalyst, the conversion became evident particularly
above 325 C.
The type of catalyst may inuence the mechanism of conversion
of triglycerides to hydrocarbons. This was demonstrated by Boda
et al. [180] who used tricaprylin and caprylic acid over Pd/AC and
NiMo/Al2 O3 catalysts. The reactions proceeded stepwise beginning
with hydrogenolysis to carboxylic acid and propane, followed by
the HDO of the former. The HDO route comprised both decarbonylation and carboxyl group reduction. The former was the dominant
reaction over Pd/AC giving C7 alkane and CO, whereas the over
the NiMo/Al2 O3 catalysts, the HDO involved the consecutive H2
addition and dehydration giving predominantly C8 alkenes, alkanes
and water. Decarbonylation proceeded via formic acid intermediate
which rapidly decomposed to CO and H2 O.
5.1.2.2. Mechanism of fatty acids conversion. Several observations
made during the conversion of free fatty acids to hydrocarbons
are common with those of carboxylic acids and esters discussed
above as part of Schemes 14. One of such study was the HPR of
mono-unsaturated oleic acid, methyl oleate and linoleic acid, carried out over Pd/C catalyst (1.52.7 MPa and 300360 C) [181]. At
the low partial pressure of H2 , the formation of di-unsaturated acids

34

E. Furimsky / Catalysis Today 217 (2013) 1356

was observed. Intermolecular isomerization (double bond migration) occurred prior to double bond HYD and deoxygenation. Under
similar conditions, stearic acid, ethyl stearate, and tristearine were
almost exclusively converted to n-heptadecane with by-products
comprising isomers of heptadecenes and undecylbenzenes [182].
Ethyl stearate was transformed via stearic acid. Gas phase products included CO, CO2 , ethane and ethylene. The concentration of
unsaturated C17 hydrocarbons from ethyl stearate was signicantly
higher than in the case of stearic acid. The formation of aromatic
structures was diminished by increasing H2 pressure, however, at
the same time the conversion of stearic acid was affected. Increasing temperature increased the conversion of ethyl stearate, but at
the same time, the formation of undesired aromatics was enhanced.
Under limited availability of active surface hydrogen, as used
in the study of Immer et al. [183a], the conversion of stearic
acid (under He) in dodecane solvent over the Pd (5%)/carbon catalyst proceeded via decarboxylation giving n-heptadecane and
heptadecenes. The HYD of heptadecenes via hydrogen transfer
facilitated by dodecane solvent resulted in 98% n-heptadecane
yield. Under H2 (10 vol.%), the stearic acid decarboxylation was
about 6-fold slower in heptadecane than in dodecane. Under He,
conversion of oleic acid involved decarbonylation, while in 10% H2
via HYD to stearic acid followed by decarboxylation.
It may be anticipated that at the same temperature, the mechanism of conversion of fatty acids may be inuenced by H2 pressure.
In an extreme case such as an inert atmosphere, the conversion of
fatty acids is dominated by decarboxylation and decarbonylation.
Under these conditions, HDO reactions also occur on the account
of an in situ hydrogen. The HYD and HDO reactions dominate the
overall conversion under high H2 pressure conditions over an active
catalyst. However, to prevent carbon loss from the parent reactant,
catalyst must exhibit a high selectivity for the HYD of carboxylic
group. Otherwise, the nal oxygen elimination proceeds via decarbonylation of the corresponding aldehydes to hydrocarbons. This
was indeed conrmed by Rozmyslowicz et al. [183b] in the study
on the conversion of model compounds such as lauric acid as well
as corresponding aldehyde and alcohol over Pd/C catalyst. Thus, in
the case of lauric acid, undecane was the main hydrocarbon product from lauric acid. The formation of dodecane would require a
catalyst with a high activity for HYD of carboxylic group and/or a
higher H2 pressure.
5.1.3. Reactants in algae biofeeds
Polyaminoacids (Fig. 1) such as polypeptides and proteins are
dominant constituents present in the algae strains. Molecular
weight of proteins usually exceeds that of polypeptides by several
order of magnitude. In various amounts, carbohydrates and lipids
are present as well. The method used for the production of biofeed
from algae biomass determines the composition of the former
[7376]. For example, little structural change between the algae
biofeed compared with that of the corresponding algae biomass is
anticipated in the case that biofeed was produced via the extraction with organic solvents (e.g., chloroform). On the other extreme,
signicant changes may be envisaged during the pyrolysis of algae
strains. Depending on conditions applied (e.g., temperature, pressure, solvent, catalyst, etc.), a signicant depolymerization may
also occur during hydrothermal treatments of algae biomass. In
this regard, a complex mixture of products may be obtained from
polypeptides and proteins while glycerides and free fatty acids
would arise from lipids. In their structure, the products obtained
from carbohydrates portions of algae, would approach that in the
biofeeds of a lignocellulosic origins.
The presence of nitrogen gives entirely new dimension to
the mechanism of HPR of the reactants which arose from the
conversion of polypeptides and protein portions (Fig. 1) of the
algae biomass. Among nitrogen compounds alkyl amines, pyrroles,

indoles, pyridines, piperidinne and pyrrolidine were clearly identied in the algae biofeeds. The HDN mechanisms involving
such reactants were extensively discussed on several occasions
[7881]. Rather than to reproduce the same material, only essential issues regarding the HDN mechanism of N-heterorings are
presented. Thus, HYD step is the essential requirement before any
N-heteroring can be opened suggesting that hydrogenated rings,
i.e., piperidines, pyrrolidines and indolines would be involved in
ring opening yielding alkyl amines as the nal intermediates before
nitrogen elimination as NH3 , takes place. In this regard, the involvement of the classical Hoffmann E eliminations and nucleophilic
SN substitutions have been conrmed [80,81]. These reactions are
initiated by proton donated by the Bronsted acidic sites, whereas
required base is supplied either by the surface of catalysts (e.g., SH
and S2 ) or NH3 . The nal elimination of NH3 may leave behind an
olen which is rapidly converted to corresponding alkane under
typical HPR conditions.
Amides as additional N-containing species are always present
in the algae biofeeds. Obviously, the mechanism of conversion of
such reactants would involve a rapid HYD of the C O entity which
is part of the amide group, followed by the hydrogenolysis of C N
bond yielding an aldehyde and amine according to the following
simplied mechanism:
The conversion of aldehydes and amines to hydrocarbons would
proceed according to established mechanisms [38,7881]. Apparently, experimental results on the mechanism of model amides
reactants under typical HPR conditions are scarce. Because of a
complex nature of such reactants, the information on their HPR
mechanism is desirable.
The total content of nitrogen in biofeeds derived from algae
biomass may approach 10 wt.%. Apparently, there is little experience in HPR of the feeds containing such high quantities of nitrogen.
This indicates on a signicant inhibiting effects of N-bases on other
HPR reactions occurring simultaneously. The extent of these effects
may vary from reaction to reaction. Inevitably, NH3 product and Ncontaining intermediates will modify the overall mechanism of the
HPR of biofeeds of algae origin. Compared with algae biofeeds, these
effects are well documented for the feeds of petroleum origin, CDL
and oil shale.

5.1.4. FTS reactants


The straight chain alkanes and alkenes are dominant components in the FTS products. The conversion of such components to
i-alkanes and i-alkenes is desirable to attain the properties of commercial fuels. For this purpose, bifunctional catalysts exhibiting a
high HCR-HIS activities, have been used [59]. Straight chain hydrocarbons are dominant components in the bio liquids of vegetable
oil origin. This suggests that there might be some similarities in the
mechanism of HPR of FTS products and those of the vegetable oil
origin.
The study conducted by Allain et al. [184] paid a specic attention to the mechanism of HCR-HIS reactions involving straight
chain hydrocarbons. In this case, the HIS-HCR activity of several
Pt catalysts supported on acidic supports was compared using
n-hexane as the reactant. The experimental observations were
interpreted in terms of the mechanism shown in Fig. 20 [184],
where C and O denote alkanes and olens, respectively. In this
scheme, the deHYD/HYD steps 1 and 5 occur on Pt metals, whereas
acidic support supplied proton for step 2. The transformation of
carbocation via step 3 was assumed to be the rate limiting step.
Consequently, the transformation of n-hexane to methyl-pentane
and methyl butanes proceeded gradually, i.e.,
nC6  2-MP  2, 3-DMB  2, 2-DMB

E. Furimsky / Catalysis Today 217 (2013) 1356

35

Fig. 20. Mechanism of hydroisomerization of n-hexane.

In general terms, a similar mechanism may apply during the


HCR-HIS of straight chain hydrocarbons with a higher carbon number than that of hexane. In fact, the reactivity of the straight
chain hydrocarbons during HCR-HIS increases with increasing carbon number of hydrocarbons [59]. In this regard, the acidity of
bifunctional catalysts (supplied by acidic support) must be properly balanced to maintain the HCR at an optimal level and to ensure
a high level of HIS.
5.2. Kinetics of HPR of bio-feeds
The objective of kinetic studies is to quantify rates of reactions
occurring during HPR. For this purpose, some mechanisms must
be simplied and/or rearranged into reaction networks with the
emphasis on the reaction steps for which the concentrations of
intermediates and products can be determined. Kinetic data can
either conrm the validity of proposed mechanisms or indicate a
need for modications. Growing interests in kinetics of HPR reactions involving model compounds such as present in bio-feeds have
been noted. Real bio-feeds have also been included in kinetic studies. For real biofeeds, additional studies may be necessary to ll the
gap in kinetic data.
5.2.1. Model compounds
Signicant advancements in kinetics of the model compounds
which are representative of those present in bio-feeds derived by
pyrolysis and liquefaction of ligno-cellulosic biomass, were made
by Laurent et al. [133,185,186] and Centeno et al. [187]. The experiments were conducted in a batch reactor at 7 MPa of H2 between
523 and 573 K. In these studies, 4-methylacetophenone (4MA),
di-ethyldecanedioate (DEC) and guaiacol (GUA), were used. Experimental results were tted using a simple kinetic equation such as,

Fig. 21. Pseudo rst order plot of conversion of 4-MA (), DES () and GUA ().
Ref. [186].
Table 11
Pseudo rst rate constants for disappearance of 4-MA, DES and GUA; % of DEC
decarboxylation; GUA phenol/cat, wt.% of coke.
Catalyst

-Al2 O3
CoMo/Al2 O3
CoMo/SiO2
CoMo/C
CoMoS

Rate constant (cm3 /min g cat)


k4MA

kDEC

% Decarbox.

kGUA

Ph/cata

wt.% coke

0.15
9.69
1.97
7.79
0.82

0.32
0.70
0.17
0.83
0.77

Nil
36
Nil
22
Nil

0.35
1.30
0.28
0.22
0.39

0
12.6
2.0
89.3
8.0

10.4
9.0
2.7

2.9

Ref. [187].
a
Ph/cat phenol/catechol ratio from HDO of GUA.

structure of the reactant. These authors reported that from kinetic


point of view, hydrogenolysis, methyl group transfer, HDO and
HYD were all important reactions occurring during the anisole conversion over Pt/Al2 O3 in the presence of H2 . The overall anisole
conversion was well represented by pseudo rst-order kinetics
with the rate constant for the disappearance of anisole being 19 L/(g
of catalyst h). The formation of some of the products could also
be well represented by rst-order kinetics, e.g., the methyl group
transfer that resulted in the formation of 2-methylphenol from

ln Xi = kWt
where Xi is the ratio of the concentration at time t to the initial reactant concentration. The pseudo-rst-order plots of these
HDO reactions are shown in Fig. 21 [185,186]. For GUA, the fast and
slow rate regions resulted from a rapid coke deposition on catalyst
surface. For DEC, a deviation from the rst-order kinetics at higher
conversions indicates interference of secondary reactions and coke.
The rate constants and the amount of coke deposit observed for
several catalysts are summarized in Table 11 [187]. A potential of
the CoMo/C catalyst should be noted, particularly because of a negligible amount of coke deposited on catalyst surface. Apparently,
an additional effort is necessary to obtain a reliable data on the
kinetics of HDO of DEC. The HDO reactions were inhibited by NH3 ,
whereas the inhibiting effect of H2 O and H2 S was less evident [186].
In the mixture, little inhibition by phenol during the HDO of methyl
heptanoate was in a contrast with a large inhibition of the HDO of
phenol by the latter [170].
The summary of observations (Fig. 22) made by Runnebaum
et al. [153] on kinetics of the conversion of anisole is introduced
to indicate a signicant complexity in spite of rather simple

Fig. 22. Kinetic routes and rate constants during conversion of anisole.
Ref. [153].

36

E. Furimsky / Catalysis Today 217 (2013) 1356

anisole. Among the primary intermediates, the highest rate of formation was observed for phenol formed directly from anisole. This
rate was more than four times higher than that of 2-methyl phenol.
The rate of 4-methyl phenol formation was much lower compared
with that of 2-methyl phenol. Formation of secondary products
such as benzene and cyclohexanone tted rst-order kinetics as
well.
Kinetics of the conversion of anisole were compared with those
of the 4-methyl anisole under identical conditions [188]. Mild
experimental conditions ensured that a large number of intermediates and primary products could be identied and quantied. It
was evident that the conversion of these reactants proceeded via
the same reaction routes (e.g., transalkylation, HDO, hydrogenolysis and HYD). Thus, for 4-methyl anisole, the products included
4-methyl phenol, 2,4-dimethylphenol, 2,4-dimetylanisole and 4methylcyclohexanone compared with phenol, 2-methylphenol,
2-methylanisole and cyclohexanone, obtained for anisole. Over
Pt/Al2 O3 catalyst, anisole was more reactive than 4-methyl anisole
with rate constants of the formation of benzene and toluene being
0.86 and 0.76 L/g h, respectively. Intramolecular transalkylations
were faster for anisole while intermolecular reactions were more
favorable for 4-methyl anisole. With respect to the overall conversion, Pt/SiO2 Al2 O3 catalyst was three times more active than
Pt/Al2 O3 catalyst. This was attributed to a higher acidity of the
SiO2 Al2 O3 compared with Al2 O3 .
Kinetic parameters were also estimated for HDO of several
other model compounds. For example, Hurff and Klein [189]
determined pseudo-rst-order rate constants for the HDO of
anisole (523593 K; CoMo/Al2 O3 ) and obtained activation energy
of 124 kJ/mol. Based on a pseudo rate constants, GUA was about 30
times more reactive than anisole.
The HPR guaiacol over Pt/Al2 O3 was investigated at 573 K and
140 kPa [153]. Dozens of reaction products were identied, with
the most abundant being phenol, catechol, and 3-methylcatechol.
The kinetically signicant reactions included HDO, HYD and
transalkylation. The HDO was evident by the production of anisole
and phenol. The HDO selectivity increased with an increasing
H2 pressure and a decreasing temperature. Products formed by
transalkylation reactions matched those produced over HY zeolite. In this case, no deoxygenated products were observed. More
than 40 products were identied during the conversion of anisole
over Pt/Al2 O3 catalyst (573 K;14 MPa) with phenol, 2-methyl phenol, benzene, 4-methyl phenol, 2,6-dimethyl phenol and 2-methyl
anisole [153].
Seven model compounds such as 4-Meguaiacol, 4-Mecatechol,
eugenol, vaniline, O-O biphenol, o-hydroxydiphenylmethane and
phenylether were used in another study on kinetics of HDO conducted over sulded CoMo/Al2 O3 , between 553613 K and at 7 MPa
by Petrocelli and Klein [190]. In this case, kinetic network was
developed for every reactant and pseudo rst order rate constants
determined for several steps in the overall HDO networks. The
effect of catalyst type on kinetics of the HDO of diphenylether was
studied by Shabtai et al. [191], while kinetics of HDO of dinaphthyl
ether were investigated by Artok et al. [192] and Kirby et al. [193].
Kinetics of the conversion of model compounds typical of those
present in vegetable oils are rather complex because of complex
HPR mechanism. For example, during the conversion of glycerides,
numerous intermediates and products (e.g., carboxylic acids, alcohols, aldehydes, alkanes, alkenes, CO2 , CO, CH4 , H2 O, etc.), are
formed. According to mechanisms shown in Figs. 18 and 19, a
series of consecutive reactions such as hydrogenolysis, decarboxylation, decarbonylation, HYD, dehydration, HDO and HIS, may
be involved. Parallel reactions, such as WGS or methanation of
CO can also occur. In simplied form, several consecutive steps
involving reactions (R1)(R5) shown in Fig. 23 were part of the
kinetic analysis conducted by Boda et al. [180]. Table 12 shows the

Fig. 23. Simplied kinetic network for conversion of triglyceride and carboxylic
acid.

product distribution during the conversion of caprilic acid. It is


evident that reaction pathways were inuenced by the type of catalyst. For example, the main product over the NiMo/Al2 O3 was H2 O,
whereas over Pd/C catalyst signicant increase in CO production
was noted. Moreover, the chain length of main hydrocarbon is one
carbon atom shorter over the latter catalyst. The appearance of
octanal, octanol and water in the product mixture (Table 12) suggests that the HDO of carboxylic acid proceeded in the consecutive
hydrogen addition and dehydration steps via reactions (R3)(R5).
Over catalyst A, alkene intermediate appeared in signicant concentration, suggesting that the reaction (R5) in Fig. 23 was the
rate-determining step of the consecutive alkane formation. For all
catalysts, octanol concentration was greater than that of octanal.
This suggests that the rate of reaction (R4) consuming octanal relative to the rate of reaction (R3) producing octanal, was greater.
The larger concentration of octanol than that of octanal also indicates that octanol conversion to octene was less favored than the
conversion of octanal to octanol and/or octene. The LM equation
was used to describe the kinetics of conversion assuming a surface
reaction as rate controlling. Thus, by disregarding the adsorption

Table 12
Yield and distribution of products during conversion of caprylic acid (CA) over
NiMo/Al2 O3 and Pd/C catalysts (360 C; 2.1 MPa total pressure).
Product (mol%)

n-C7
i-C8
n-C8
C8
n-C8
Hydrocarbon
Octanal
Octanol
CA
H2 O
CO
CH4
CO2
CA conversion
C7 /(C7 + C8 )

Catalyst
Aa

Bb

Cc

Dd

Ee

2.0
1.0
12.5
5.6
19.1
21.1
0.7
1.5
31.7
43.0
2.0
0.0
0.0
68.3
0.09

7.5
0.6
12.4
1.2
14.2
21.7
0.8
2.6
29.8
39.7
4.7
0.5
0.2
70.2
0.34

1.9
4.8
0.7
0.8
6.3
8.2
2.4
3.7
63.1
20.5
2.0
0.0
0.3
36.9
0.23

15.5
0.2
0.8
0.2
1.2
16.7
0.2
0.5
52.5
17.9
12.2
0.0
0.0
47.5
0.93

20.3
0.2
0.5
0.2
0.9
21.2
0.4
0.7
40.8
20.1
16.7
0.0
0.2
59.2
0.96

Ref. [180].
a
9Mo2.5Ni(P, Si) commercial catalyst.
b
10Mo3.3Ni (P).
c
11Mo3.0Ni laboratory made catalyst.
d
3Pd/C.
e
5Pd/C commercial catalyst.

E. Furimsky / Catalysis Today 217 (2013) 1356

37

Table 13
Rate constants (min1 ) for decarboxylation of ethyl stearate.
Rate constant
6.27 1012
1.31a 103
1.01 103
1.45a 1012
2.47a 103
2.31a 104
4.55 104

k1
k2
k3
k4
k5
k6
k7

Ref. [195a].
a
Fixed values based on modeling of stearic acid.

Fig. 24. Kinetic network for decarboxylation of ethyl stearate.

of products on active sites, the following rate equation could be


obtained:
r=

kKH2 pH2 KCA pCA


(1 + KH2 pH2 + KCA pCA )2

(1)

where k is rate constant, KH2 and KCA are the equilibrium adsorption
coefcients. as well as pH2 and pCA partial pressure of H2 and caprilic
acid, respectively. In the case that 1 + KCA pCA term is much greater
than (1 + KH2 pH2 ) term, the following simplied form of the LM
equation is obtained:
r=

kKH2 pH2
KCA pCA

(2)

At high pressure of H2 , i.e., 1 + KCA pCA  KH2 pH2 , the following form
of the L-M equation is obtained:
r=

kKCA pCA
KH2 pH2

(3)

Boda et al. [180] showed that dr/dpCA , obtained from Eq. (1), was
positive if KCA pCA < 0.5 and negative if KCA pCA > 0.5. Similarly, the
dr/dpH2 was positive if KH2 pH2 < 0.5 and negative if it is >0.5.
Resasco et al. [194] studied kinetics of the conversion of methyl
stearate in tetradecane solvent in the semi-batch reactor between
573 and 623 K, 0.69 MPa over Pt(1%)/Al2 O3 . Kinetics could be
described by a general rate law kinetic expression such as
r=

dCms
n
= k eE/RT Cms
dt

The values of kinetic parameters such activation energy, rate constant and reaction order, estimated from the t of experimental
data were 107 kJ/mol, 2.1 104 and 0.31 s1 , respectively.
The detailed kinetic study on the transformation of ethyl
stearate was conducted by Snare et al. [195a]. The experiments
were carried out in a semi-batch reactor at 270360 C after the
commercial 5% Pd/C catalyst was reduced in the ow of H2 at 200 C
for 2 h. The powder form of the catalyst ensured the absence of
diffusion limitations. The reaction medium included n-dodecane
solvent as well as either an inert gas (e.g., He or Ar) or Ar + 5 vol.%
H2 . The pressure varied from 1.7 to 4.0 MPa. Rather mild conditions were suitable for quantication of intermediates and products
concentrations to be incorporated in the kinetic network shown in
Fig. 24 [195a]. The kinetic modeling was based on the LH law.
The adsorption constants of gaseous products were not included
in the denominator because of very low values. Surface reaction
was assumed to be rate limiting, while the adsorption steps were
assumed to be rapid. Based on the network in Fig. 24, total of seven
rate expressions were developed. The following is an example of
the rate expression for reaction step 1:
r1 =

k1 KA cA
1 + KA cA + KB cB + KC cC + KD cD + KE cE

where k1 is the reaction constant for step 1, while Kx and cx are


adsorption constants and concentrations of the species in the network shown in Fig. 24. A system of differential equations for every
species in kinetic network could be derived using mass balance
and stoichiometry of the reactions. For example, for the step 1, the
following expression was obtained:
1 dcA
= r1 r3 r7
 dt
where  denotes the catalyst bulk density. Based on this analysis,
the apparent activation energy between 300 and 360 C approached
57 kJ/mol, while the reaction order of stearic acid conversion
approached zero because of a rapid initial catalyst deactivation. The
apparent rate constants at 300 C are listed in Table 13 [195a]. Initially, the conversion of ethyl stearate was dominated by reactions
3 and 7 in Fig. 24. The study of Snare et al. [195a] was not conducted under typical HPR conditions. However, it is incorporated
in this review because of the level of details relevant for the kinetics
of decarboxylation of alkyl esters.
The studies conducted under non-conventional HPR conditions
such as either in subcritical or supercritical water, i.e., below and
above 370 C, respectively, have been receiving attention. The reaction network on the HDO of benzofuran established by Dickinson
et al. [195b] after an extensive characterization of products was
used for determination of kinetic parameters. The hydrogen to
reactant ratios, i.e., 0:1, 0.5:1, 2:1, 3:1, 4:1, and 6:1 were chosen to be moderately substoichiometric with respect to producing
ethylcyclohexane and the stoichiometric ratio 6:1. The experiments
were performed in an autoclave over Pt(5%)/C catalyst at 380 C.
A self-inhibition by benzofuran more evident than inhibition by
intermediates such as ethylphenol and ethylcyclohexanol. Therefore, the 1 + KBF CBF term was incorporated in the denominators
of the LM kinetic model to account for this inhibition. In addition, several rate equations for the major reaction pathways were
derived. The experimentally determined kinetic parameters agreed
with those obtained from the kinetic model.
5.2.2. Real bio-feeds
The following tentative scheme depicts the overall conversion
of biocrude to biofuel:
Gas+water
Gas

Bio-crude

Bio-feed

Coke
Stabilization

Bio-fuel
Coke
Upgrading

This scheme would be more applicable to bio-crudes from the


biomass of a lignocellulosic origin. For these biofeeds, the stabilization step would be much more desirable for pyrolysis liquids
compared with liquids obtained by biomass liquefaction [196].
According to the results published by French et al. [197], the stabilization step is usually conducted under rather mild HPR conditions,
followed by the nal upgrading step carried out under more severe

38

E. Furimsky / Catalysis Today 217 (2013) 1356

HPR conditions. A similar stabilization step may not be necessary


for the bio-crudes of vegetable oils origin.
Based on the above rational, kinetic studies paying attention to
both the stabilization and upgrading stages are of a primary interest. The objective of the stabilization step is to maximize the rate of
bio-feed production. In this regard, temperature was key parameter
as it was reported by Rocha et al. [198]. Thus, signicant increase
in gas formation was observed above 623 K. The yield of bio-feed
could be further increased by increasing H2 pressure. High rates
of HDO and HYD in the second stage conducted over NiMo/Al2 O3
ensured low aromaticity and oxygen content products.
Sawdust and bark pyrolysis oils were used as the feeds for HPR
in a trickle bed reactor. The objective was a single stage kinetics of
HDO. Changes in the product composition as a function of reaction
temperature, H2 pressure, LHSV and catalyst type, were investigated [172,199,200]. Several catalysts, i.e., Pt/Al2 O3 SiO2 , sulded
CoMo/Al2 O3 , NiW/Al2 O3 and NiMo/Al2 O3 , were used. The reaction temperature and pressure ranged from 623 K to 673 K and 5.2
to 10.4 MPa, respectively. Two kinetic models, one for the overall HDO and the other for the compositional changes of products,
were developed. The reaction order was inuenced by the type of
catalyst and it was 1.0 for the Pt/Al2 O3 SiO2 and 0.3 for the other
catalysts. The overall HDO was not a function of LHSV but could
be modeled by an empirical function of temperature and pressure, while pseudo rst order reaction network was used to relate
products composition using ve lumps model.
A similar kinetic expression as used by Sheu et al. [200] was used
in the study conducted by Su-Ping et al. [201], i.e.,
dCox
m
= k Cox
Pn
dt
where Cox , k and P are concentration of oxygen in products at
time t, rate constant and H2 pressure, respectively, whereas m and
n are corresponding reaction orders. Because of low H2 pressure
dependence (from 1.5 to 3.0 MPa), this equation was simplied
by incorporating H2 pressure into the rate constant k to give the
following form of kinetic expression:
dCox
m
= k Cox
dt
The experimentally determined reaction order of 2.3 differed from
1.0 assumed in the study of Sheu et al. [200]. This may be attributed
to different experimental set-ups, i.e., continuous vs batch reactors
and different composition of biofeeds. As pointed out by Mortensen
et al. [202], the difference in reactivity of the reactants in the
feed may be another contributor. Thus, highly reactive species are
removed at high rates before the rate rapidly declined due to very
low reactivity of reactants left behind.
Kinetics of the HDO of the triglycerides were studied by Sebos
et al. [203] in plug ow xed bed reactor at 593 K, 3 MPa and H2 S
partial pressure of 1 vol.%. It was assumed that the rate HDO of ester
groups was rst-order reaction, proportional to ester concentration
in the feed, i.e.,
rHDO = kHDO Cest0 A (1 x),
where rHDO is the rate of HDO, kHDO the reaction rate constant,
Cest0 the ester group content of the feedstock (determined by
saponication number) and x their conversion. By rearranging this
equation using the mass balance of ester group content, the following equation was obtained:
kHDO =

ln(1 x) mfr
mcat

where mfr is the mass feed rate and mcat is the mass of catalyst. At
WHSV less than 5 h1 a conversion of almost 100% was be achieved.
The activation energy of HDO approached 111 kJ/mol.

Maki-Arvela et al. [204] studied the effect of temperature


(300350 C), initial concentration of feed (0.150.6 mol/L) and
metal loading (04 wt.%) on the kinetics of conversion of tall oil
fatty acids over 1 wt.% Pd/C. The total conversion of fatty acids
increased with increasing temperatures and decreasing initial feed
concentrations giving n-heptadecane and n-heptadecene as the
main products. In addition, linear C17 hydrocarbons and aromatic
C17 compounds (e.g., undecylbenzene) were formed. An increase of
the metal loading (4 wt.%) led to an increase of the selectivity for
linear C17 hydrocarbons.
Spent vegetable oil was used as the feed to study kinetics of the
conversion of the long residue, dened as 375 FBP C fraction, in
a batch microreactor over temperature range of 400430 C, reaction time of 3090 min and initial H2 pressure of 1.03.0 MPa over
sulfated zirconia [205a]. Based on the experimental data, the following linear expression for the production of gasoline fraction,
was obtained:
gasoline yield = 16.47 + 1.46x1 + 1.405x2
where x1 is the temperature and x2 is the reaction time. Two kinetic
expressions, one for the 1st order and the other for 2nd order kinetics were derived, e.g.,
ln(CB ) = ln(CBo ) k1 t (1st order)
1
1
=
+ k2 t
CB
CBo
where CBo and CB are concentration of long residues in the feed and
products, respectively. The plot of experimental data gave a straight
line for the 2nd order kinetics. In this case, the kinetic model was
dened as k(s1 ) = 914.886 exp[(83.439 kJ/mol)/(RT)].
From rening point of view, the information on kinetics of the
blends of petroleum feeds with bio-feeds, is of interest. In this
regard, important studies were conducted by Huber et al. [168]
and Donnis et al. [172]. Using the blend of sun ower oil and VGO
derived from a petroleum, the former authors found that the rate of
oxygen removal from sunower oil was much faster than the HDS
of VGO. Thus, at 300 C over NiMo/Al2 O3 catalyst, the conversion of
sunower oil was 100% while sulfur removal was only 41%. However, sunower oil in the concentration up to 50% did not inhibit the
rate of HDS. Similar observation was made during the HPR of the
blend of rapeseed oil with LGO over the sulded NiMo/Al2 O3 catalyst, while for the same blend, the selectivity for decarboxylation
and the HYD of aromatics was decreased [172].
A jatropha biofeed was used to study kinetics of the HPR of
triglycerides over the NiMo and CoMo catalysts supported on mesoporous titanosilicates at 320360 C, 28 MPa and 48 h1 [205b,c].
The kinetic analysis was based on quantication of four lumps, i.e.,
C5 C8 , C9 C14 , C15 C18 and >C18 with their concentration designation as CL , Cj , CH and Cp and corresponding rate constants as
k1 , k2 , k3 and k4 , respectively. The following set of equations was
developed:
Ctg

= ek t
Ctgo
CL =

k1

[ek t 1]
k Ctgo

Cj =

Ch =

k2

[ek t 1]
k Ctgo

k3

[ek t 1]
k Ctgo

E. Furimsky / Catalysis Today 217 (2013) 1356


Table 14
Rate constant for kinetics of HPR of triglycerids [205b].

k
k1
k2
k3
k4

Cp =

300 C

320 C

7.4
0.03
0.02
0.07
7.3

14.35
0.04
0.11
1.24
13.25

k4

[ek t 1]
k Ctgo

39

in most of biofeeds, a sulfur donating agent must be added to


ensure stability of active phase in conventional catalysts. Numerous
attempts to develop novel catalysts for HPR of biofeeds have been
focusing on catalysts not requiring presulding and/or the presence of sulfur in biofeed while still exhibiting a high activity and/or
resistance to deactivation. In this case, noble metals supported on
various supports as well as metal carbides, nitrides and phosphides
of transition metals, have been receiving most of the attention. This
will be illustrated on several examples presented in the following
sub-chapters.
6.1. Catalysts for HPR of biofeeds of ligno-cellulosic origin

where k = k1 + k2 + k3 + k4 . The rate constants estimated at 300 and


320 C from the yield of lumps are shown in Table 14 [205b].
It is evident that at these temperatures, the formation of heavy
products (>C18 ) was far predominant. However, the yield of such
products decreased with further temperature increase while that of
the C9 C14 and C15 C18 fractions increased. Activation energies of
83, 126, 47 and 47 kJ/mol were obtained for naphtha, middle distillates, heavy distillates and oligomerized products, respectively
over sulded CoMo/Al2 O3 catalyst [205c]. Above 6 MPa, the yield of
(C15 C18 ) lump increased while that the >C18 products decreased.
Database of kinetic parameters determined in these studies is perhaps one of its kind in the literature. In the study of Kovacs et al.
[205d], further improvement in triglycerides conversion to hydrocarbons can be obtained above 360 C between 2 and 4 MPa over
uorinated NiMo/Al2 O3 catalyst.
6. Catalyst development for HPR of biofeeds
Various catalyst formulations consisting of the conventional
metals such as Co(Ni)/Mo(W) either in an unsupported form or in
combination with a wide range of different supports, have been
tested for the HPR of biofeeds. Because of the absence of sulfur

The upgrading of primary liquids derived from lignocellulosic


biomass via HPR was compared with cracking option by Hicks
[206]. It was evident that the upgrading via the former option may
require large amount of hydrogen and catalyst inventory. Therefore, catalyst development for these application is essential to make
HPR of biofeeds viable a option. As the summary of studies in
Table 15 shows, more attention has been paid to non-conventional
catalysts.
6.1.1. Conventional catalyst
Both unsupported and supported catalysts have been used for
the HPR of biofeeds of lignocellulosic origin. Catalyst development
and testing usually considered two stages, i.e., the rst stage, conducted at moderate temperatures (e.g., 170270 C; 13 MPa of H2 )
and the second stage, conducted at a similar H2 pressure but at
much higher temperatures (e.g., 400 C) [206a,207]. The objective
of the rst stage was the removal of thermally unstable components
(e.g., guaiacols, biphenols, etc.) with the aim to extend catalyst
life in the second stage. Under the latter conditions, lignin alone
or lignin in biofeeds can be converted to phenols and hydrocarbons such as cyclohexane, benzene, naphthalene, phenanthrene,
etc. [208211].

Table 15
Catalysts for HPR of biofeeds derived from lignocellulosic biomass.
Conventional catalysts

Feed

Conditions

Ref.

1. Unsupported and supported sulded Co(Ni)Mo(W)


2. Unsupported and supported MoS2
3. Unsupported MoS2 and CoMoS
4. Sulded Co(Ni)Mo/Al2 O3
5. CoMoW supported on SBA-15 and SBA-16
6. CoMoS supported on ZrO2 , -Al2 O3 and TiO2
7. MoS2 on activated carbon

Lignin
Guaiacol
Phenol
PLCB
Anisole
Anisole and guaiacol
Guaiacol

Autoc.; 270 and 400 C, 13 MPa


Cont.; 300 C K; 4 MPa
Batch.; 350 C; 2.8 MPa
Cont.; 400 C, 13.7 MPa
Cont.; 310 C; 3 MPa
Cont.; 300 C; 4 MPa
Batch; 300 C; 5 MPa

[207]
[212,213]
[213b]
[206b]
[152]
[219a]
[219b]

Pyrolysis oil
Anisole
Anisole

Cont., 170 C; 13.7 MPa


Cont; 400 C; 1 atm
Cont; 250400 C; 0.52.0 MPa

[206b]
[220]
[221a]

Guaiacol
Guaiacol

Cont; 300 C; near atm.


Batch; 330 C; 8 MPa

[149,150a,b]
[151,224]

Guaiacol

Batch; 250 C; 4 MPa

[222]

Non-conventional catalysts
1. Sulded Ru/carbon
2. Bifunctional Pt/H
3. Pt, Pd and Rh combined with oxides of Mo, W, Co,
Mn, Zr, Ce, Y, Sr and La
4. Pt supported on Al2 O3 , H zeolite and MgO
5. Rh, Pt, Pd, RhPt, RhPd and PdPt supported on
ZrO2
6. Bifunctional Pt, Rh, Pd and Ru catalysts supported on
Al2 O3 , SiO2 Al2 O3
7. Pt, Sn and PtSn on inconel monoliths coated with
CNF
8. Pt, Pd and PtPd supported on SiO2 Al2 O3
9. Ru/C, Ru/TiO2 , Ru/Al2 O3 , Pt/C and Pd/C
10. Rh, Pt and Pd supported on ZrO2
11. Re/ZrO2 -sulfated and Re/ZrO2
12. Ni2 P/SiO2 , MoP/SiO2 and NiMoP/SiO2
13. Ni2 P, Fe2 P, MoP, Co2 P and WP all supported on SiO2
14. Mo2 N/C
15. Mo2 N supported on Al2 O3 and SBA-15
16. Co(Ni)Mo(W) borides
17. Ni-Cu combined with -Al2 O3 , SiO2 , SiO2 ZrO2 ,
La2 O3 and CeO2 ZrO2
18. Ni/Al2 O3 doped with In2 O3

Guaiacol and anisole

Cont.; 400 C, atm. N2 + H2

[223]

Benzofuran
Pyrolysis oil
Pyrolysis oil
Guaiacol
Anisole
Guaiacol
Guaiacol
Guaiacol
Phenol, acetophenone, benzaldehyde
Guaiacol

Cont; 280 C; 3 MPa


Batch; 350 C; 20 MPa
Batch; 350 C; 20 MPa
Batch; 300 C; 5 MPa
Cont.; 300 C; 1.5 MPa
Cont.; 300 C; 1.5 MPa
Batch; 300 C; 5 MPa
Batch, 330 C; 5 MPa
498 or 523 K, 4 MPa
Batch; 320 C; 17 MPa

[224b]
[225]
[228]
[229]
[230]
[141]
[231a,232a]
[231b,232b]
[214218]
[233a]

C2 C4 oxygenates

Batch; 220380 C; 2.1 MPa

[233b]

40

E. Furimsky / Catalysis Today 217 (2013) 1356

Table 16
Conversion of GUA (mol g1 s1 ) and intrinsic HDO rates (mole Mo1 s1 ) at 573 K.
Catalysts

MoS2
CoMoS
MoS2 /Al2 O3
CoMoS/Al2 O3

Conversion

Intrinsic rates

Guaiacol

HDO

Guaiacol

HDO

35
222
64
71

10.1
49.7
2.3
5

6
72
81
93

1.7
16.2
2.9
6.6

Ref. [213a].

6.1.1.1. Unsupported catalysts. Bulk MoS2 prepared by decomposition of ammonium thiomolybdate and the bulk CoMoS catalysts
prepared in aqueous or mixed solution before being sulded (15%
H2 S in H2 ) were used for the HDO of guaiacol [212,213]. The addition of Co promoter increased guaiacol conversion to phenol via
demethoxylation. Phenol was converted to benzene via direct HDO.
Table 16 [213a] shows much greater activity of the unsupported
catalysts compared with their supported counterparts. Yoosuk et al.
[213b] reported that for the unsupported MoS2 and CoMoS2 catalysts, the activity and selectivity during the HDO of phenol could be
controlled by the conditions of catalyst preparation. The addition of
Co to MoS2 signicantly enhanced activity. The amorphous (unsupported and unsulded) Co(Ni)Mo(W)B catalytic phases studied by
Wang et al. [214218] are discussed in more details in a separate
section.
6.1.1.2. Supported catalysts. Different supports and/or a modied
-Al2 O3 support may improve performance of the catalysts containing conventional metals compared with the corresponding
catalysts supported on the unmodied -Al2 O3 . This was evident in
the study of Laurent et al. [133] on the HDO of guaiacol. For example,
the carbon support was superior to that of the -Al2 O3 (Fig. 9).
To evaluate other supports, the phosphate-modied SBA-15 and
SBA-16 materials were prepared via incipient wetness impregnation with aqueous solutions of H3 PO4 to obtain substrates with
P2 O5 loadings of 0.5 and 1.0 wt.% [152]. After water evaporation
at room temperature, the solid was dried at 110 C for 4 h and
then calcined at 500 C for 4 h. The CoMoW catalysts supported
on these materials were prepared by incipient wetness impregnation. The catalysts were used for hydrogenolysis of anisole at
310 C and 3 MPa. The sulded CoMoW/SBA-16 catalyst containing 0.5 wt.% of P2 O5 exhibited the highest activity and stability. The
most active catalyst had the highest total acidity, the largest population of Mo(W)S2 slabs and the greatest suldation degree of the
W species.
Study on the effect of support on the catalyst activity at 310 C
and 3 MPa during the HDO of anisole and guaiacol revealed a
remarkable promoting effect of ZrO2 compared with -Al2 O3 and
TiO2 [219a]. Moreover, the CoMoS/ZrO2 catalyst was very selective toward CAR O bond hydrogenolysis which followed after
demethoxylation. Over -Al2 O3 supported catalysts, the presence
of methylation reactions was evidenced by the formation of heavier
products and enhanced formation of coke.

Two activated carbons (AC) with different porosity and concentration of oxygen surface groups were used as support for MoS2 for
the HDO of guaiacol (batch reactor at 300 C and 5 MPa) [219b]. The
oxygen surface groups (e.g., mainly carboxylic, quinonic and lactonic) affected the dispersion of Mo species on the support surface.
As the consequence, activity of the MoS2 /AC catalysts supported on
the preoxidized supports was lower.
6.1.2. Unconventional catalysts
Among non-conventional catalysts, noble metals (e.g., Pt, Pd, Rh,
etc.) supported on various supports as well as the carbides, nitrides,
phosphides and borides of transition metals (Table 15), have been
tested.
6.1.2.1. Noble metals containing catalysts. Both, transalkylation and
HDO reactions were observed by Zhu et al. [220] during the HDO
of anisole (400 C and 1 atm. H2 ) over bifunctional Pt/H catalyst.
Thus, benzene, toluene and xylenes were present among the products. In contrast, over monofunctional catalysts, such as H and
Pt/SiO2 , the rates of these reactions were much lower.
Yakovlev et al. [221a] introduced a novel concept of bifunctionality. In this case, a bifunctional catalyst should comprise an
oxide form of a transition metal to facilitate the activation of oxygroups and a noble metal in its reduced state to activate hydrogen.
The oxides of metals such as Mo, W, Co, Mn, Zr, Ce, Y, Sr and La
were used to supply the former functionality while Pt, Pd and Rh to
ensure hydrogen activation. Series of the tested catalysts included
the Rh, Rh, Co, Ni and NiCu-containing catalysts in combination SiO2 , Al2 O3 , ZiO2 , CeO2 , CeO2 ZrO2 , as supports. The tests
were conducted in the xed bed reactor at 250400 C and total
pressure of 0.52.0 MPa. The results are shown in Table 17 [221a].
The overall conversion over RhCo/Al2 O3 , RhCo/SiO2 , Rh/ZrO2 and
Rh/CeO2 catalysts approached 100% while selectivity dened as
the ratio of aliphatics/aromatics exhibited a signicant variation, i.e., 0.301.68. The NiCu/Al2 O3 catalyst having the Ni/Cu
ratio of eight exhibited high activity during the HPR of biofeed
obtained by fast pyrolysis oil [221b]. These results clearly indicate
on a signicant potential for developing a catalyst not requiring
presulding.
The HDO of guaiacol over Pt/Al2 O3 at 573 K (near atmospheric
H2 ) took place with a high selectivity for CAL O bond cleavage
[149]. The most abundant products included phenol, catechol and
3-methylcatechol. Less evident were methyl group transfer reactions. Alkyl transformation reactions were predominant over Pt/H
zeolite [150a] as well as on zeolite alone [149]. However, for the
latter, the activity for oxygen removal was very low. Transalkylation reactions were absent over the Pt/MgO catalyst [150b] yielding
phenol, catechol and cyclopentanone as the main products in
addition to methane, n-butane, butenes, n-pentane and carbon
monoxide. The Pt/MgO catalyst was less prone to deactivation than
the Pt/Al2 O3 catalyst.
Bifunctional catalysts containing Pt(5 wt.%), Rh(3 wt.%),
Pd(5 wt.%) and Ru(5 wt.%) supported on Al2 O3 as well as on
SiO2 -Al2 O3 and nitric-acid-treated carbon black were used for

Table 17
Activity of Rh-based catalysts during HDO of anisole at 300 C and 1.0 MPa of H2 .
Catalyst
1

LHSV (h )
Al/Ara
Conver.b
HDO (%)

Rh/SiO2

RhCo/Al2 O3

Rh/CoSiO3

RhCo/SiO2

Co/SiO2

Rh/ZrO2

Rh/CeO2

Al2 O3

0.3
0.22
53.4
30.4

0.3
1.68
98.0
74.7

0.5
0.81
82.0
79.0

0.3
0.30
99.0
81.0

0.3
0.03
10.3
6.3

1.0
0.88
99.6
90.8

0.4
0.30
100.0
94.6

0.3
0.01
100.0
0.5

Ref. [221a].
a
Aliphatic/aromatic molar ratio.
b
Total conversion in %.

E. Furimsky / Catalysis Today 217 (2013) 1356

the HDO of guaiacol (250 C; 4 MPa of H2 ; batch reactor) [222].


Except the Pt/SiO2 Al2 O3 and Pd/carbon black, the conversion of
guaiacol on the other catalysts approached 100% with selectivity
to cyclohexane varying from 14 to 60%. The highest conversion
and selectivity to cyclohexane was observed over Rh/SiO2 Al2 O3
and Ru/SiO2 Al2 O3 catalysts.
Gonzalez-Borja and Resasco [223] prepared an inconel monoliths coated with the in situ-grown carbon nanobers (CNF).
Coating with the CNF increased surface area and facilitated anchoring sites for active metals. The single metals such as Pt and Sn as well
as bimetallic PtSn catalysts were prepared by the impregnation
of the monoliths. The catalysts were used for the HDO of guaiacol
and anisole. The main products obtained from these model feeds
were phenol and benzene. The bimetallic PtSn catalysts exhibited
higher activity and stability than monometallic Pt and Sn catalysts.
The Rh as well as RhPt and RhPd catalysts prepared by the
incipient wetness using ZrO2 support were compared with conventional sulded CoMo and NiMo catalysts during the HDO of guaiacol
[151]. The former catalysts were more active in a mono-metallic
form. Thus, the addition of the second metals to Ru (e.g., Pt or Pd)
had no benecial effect on the HDO activity. In a similar study conducted by Gutierrez et al. [224a], the monometallic Rh, Pd, and Pt
as well as bimetallic RhPd, RhPt and PdPt catalysts supported on
ZrO2 were compared with the sulded CoMo/Al2 O3 catalyst during
the HDO of guaiacol (3 wt.% in n-hexadecane) at 300 C, 8 MPa in
a batch reactor. The activity of the noble metal catalysts was similar or better than that of the conventional sulded CoMo/Al2 O3
catalyst. However, the carbon deposition on the noble metal catalysts was much less evident compared with that on the CoMo/Al2 O3
catalyst. The Rh/ZrO2 catalyst exhibited the highest activity.
The activity and selectivity of monometallic Pt and Pd as well
as bimetallic PtPd catalysts supported on SiO2 Al2 O3 were tested
during the HDO of benzofuran (ow reactor, 280 C, 3 MPa) [224b].
The PdPt (Pd/Pt = 4) catalyst exhibited the highest activity yielding
ethylcyclohexane and methylcyclohexane as the main products.
The latter product indicated the involvement of cracking caused
by the acidic support.
Methyl octanoate in tetradecane was used by Do et al. [194]
to compare activity of the Pt(1%)/TiO2 catalyst with that of the
Pt(1%)/Al2 O3 catalyst. The former catalyst exhibited a higher selectivity to C8 hydrocarbon because of a larger oxygen vacancy
availability on the TiO2 support than that on Al2 O3 support. Under
He, formation of heavy products such as diheptyl ketone, npentadecane, and octyloctanoate, was quite evident. These species
were almost completely removed from the products under H2 .
The noble metals containing catalysts (e.g., Ru/C, Ru/TiO2 ,
Ru/Al2 O3 , Pt/C and Pd/C) were used for the HPR (350 C; 20 MPa;
batch reactor) of pyrolysis biofeed [225]. The best performance was
exhibited by Ru/C catalyst. This catalyst was used to study the effect
of reaction time (16 h) on the yield of liquid products [226]. The
highest yield was obtained at reaction time of 4 h. The trends in
hydrogen consumption complemented those in the yield of liquid products. The yield increased with contact time and reached
maximum at 4 h. Further increase in reaction time decreased yield
of liquid and increased that of gaseous products. However, the
H/C ratio of liquids exhibited a continuous increase from 1.05 to
1.35 by increasing reaction time from 1 to 6 h. The activity of the
Ru/C could be further increased by optimizing the method of catalyst preparation [227]. Thus, by using different precursors such
as RuCl3 , Ru(NO)(NO3 )3 and Ru(acac)3 , the highest activity was
obtained using the RuCl3 with Ru loading of 5 wt.%. However, a
long term performance of the Ru/C catalyst requires attention.
Thus, loss of activity after recycling used catalyst was quite evident.
For recycling, the catalyst was separated from liquids at the end
of experiment, thoroughly washed with acetone and dried before
being added to the reactor for next cycle.

41

Fig. 25. Activity of noble metals supported on ZrO2 expressed on active metal basis
(350 C; 2 MPa; 4 h; batch).
Ref. [225].

The bio oil from fast pyrolysis of biomass was used as the feed for
HPR (batch reactor, 350 C, 20 MPa total pressure, 4 h reaction time)
over mono- and bi-metallic catalysts consisting of Rh, Pt, Pd supported on ZrO2 [228]. The highest activity was exhibited by Pd/ZrO2
whereas Pt/ZrO2 was the least active. For the latter, an incomplete
reduction of Pt or a low sulfur tolerance were responsible for low
activity. However, the highest quality products were obtained over
Rh/ZrO2 . Fig. 25 [225] shows that the activity of all noble metal catalysts and/or hydrogen consumption, expressed on an active metal
basis, was higher than that of the conventional CoMo/Al2 O3 catalyst but a factor of two lower when comparing on a total catalyst
metal loadings.
Effect of sulding in either H2 /H2 S or N2 /H2 S mixture on the
activity of the Re/ZrO2 and Re/sulfated-ZrO2 catalysts during the
HDO of guaiacol were investigated by Ruiz et al. [229] in a batch
reactor at 300 C and 5 MPa. More active catalysts were obtained
after sulding in the N2 /H2 S. For both sulding mixtures, the activity of the Re/ZrO2 was higher that of the Re/sulfated-ZrO2 .
phosphides. The
Ni2 P/SiO2 ,
MoP/SiO2
and
6.1.2.2. Metal
NiMoP/SiO2 catalysts with the P/Me (Me = Mo or Ni) molar
ratio of one were prepared by Li et al. [230] and tested during the
HDO of anisole in xed-bed reactor. The HDO activity decreased
in the following order: Ni2 P/SiO2 > NiMoP/SiO2 > MoP/SiO2 . The
higher activity of Ni2 P than MoP was attributed to higher d
electron density in Ni2 P. In general, Ni phosphide catalysts were
more active than the conventional NiMo/Al2 O3 catalyst. However,
for the former catalysts, the reaction with water may lead to
the formation of metal oxide and phosphate causing catalyst
deactivation.
The HDO activity of Ni2 P/SiO2 , Fe2 P/SiO2 , MoP/SiO2 , Co2 P/SiO2
and WP/SiO2 was tested using guaiacol as a model compound [141]. The activity exhibited the following order:
Ni2 P > Co2 P > Fe2 P WP MoP. The unwanted catechol was
not observed among the products over the most active Ni2 Pd/SiO2
even at short contact time compared with the other catalysts.
The phosphides were more active than commercial CoMo/Al2 O3 .
The latter catalyst deactivated quickly due to coke deposition. For
most phosphides, the major products were benzene and phenol. A
commercial Pd(5 wt. %)/Al2 O3 catalyst was more active than the
metal phosphides but it produced only the catechol intermediate.
6.1.2.3. Metal carbides and nitrides. The carbon supported Mo2 N
catalysts were prepared by wetness impregnation of three different
activated carbons and nitrided with ammonia at 700 C [231,232].
The catalysts were used for the HDO of guaiacol (batch reactor at

42

E. Furimsky / Catalysis Today 217 (2013) 1356

5 MPa of H2 and 300 C). The HDO activity was support dependent
and could be further increased by activation in H2 for 6 h at 400 C.
For unsupported Mo2 N catalysts, the activity increased with the
increasing N/Mo ratio of the catalysts. This was indicated by the
increased conversion and the increased phenol/catechol ratio. The
addition of Co increased overall HDO of guaiacol.
Ghampson et al. [232a,b] prepared Mo nitride catalysts supported either of -Al2 O3 or on SBA-15. The nitridation to obtain Mo
nitrides was performed by thermal conversion either in ammonia
or in N2 /H2 mixtures. The catalysts were used for the HDO of guaiacol (autoclave at 300 C and H2 pressure of 5 MPa). The catalysts
prepared by nitridation in ammonia had a higher N/Mo ratio and
were more active. Over the -Al2 O3 supported Mo nitride, guaiacol conversion involved both demethylation and demethoxylation
while the latter pathway dominated the overall guaiacol conversion
over the SBA-15 supported Mo nitride. Under identical conditions,
the Mo nitride supported on -Al2 O3 was more active than conventional NiMo/Al2 O3 catalyst.
In the subsequent study conducted by Ghampson et al. [232c],
four Mo nitrides supported on different activated carbons were
prepared by impregnation with an aqueous solutions of ammonium heptamolybdate and subsequent conversion to nitride using
either NH3 or a N2 /H2 mixture. The bimetallic MoCo nitrides were
prepared by sequential impregnation rst with Mo precursors followed by that with Co nitrate after the former was dried and
calcined. The effects of preparation procedure and type of activated carbons were determined using the model reaction such as
the HDO of guaiacol under similar conditions as used in the previous study [232b]. The transformation of guaiacol proceeded via
direct demethoxylation rather than via the formation of catechol.
The catalysts prepared in the N2 /H2 mixture exhibited a higher
activity than those prepared via ammonolysis. This was attributed
to a higher dispersion of Mo oxynitride in the former catalysts.

SiO2 , SiO2 ZrO2 , La2 O3 and CeO2 ZrO2 ) and used for the HDO of
guaiacol alone, i.e., neither solvent nor sulfur donating agent were
present [233a]. The catalysts were activated directly in the batch
reactor in the ow of H2 for 1 h at 400 C. After activation, the experiments were conducted at 320 C and 17 MPa of H2 . For all catalysts,
the overall guaiacol conversion ranged from 80 to 97%. However,
Ni36.5Cu2.3/ZrO2 SiO2 La2 O3 , Ni57.9Cu7.0/SiO2 and Ni55.4/SiO2
exhibited the highest HDO activity yielding 9297% hydrocarbons.
The product distribution observed in this study [233a] was in line
with the mechanism of the overall conversion of guaiacol, generally
established in many studies [38]. During the HPR of anisole and a
fast pyrolysis oil, the bimetallic 16Ni2Cu catalyst supported on Al2 O3 was much more active than monometallic Ni and Cu analogs
[233b].
Because of a high content of C2 to C4 oxygenates (carboxylic
acids, ketones and aldehydes) in the pyrolytic oil fraction boiling below 150 C, a selective HYD to alcohols rather than HDO to
hydrocarbons may be more desirable. Thus, the alcohols are very
good blending components with gasoline. The selectivity for alcohol formation over Ni/Al2 O3 catalyst was signicantly enhanced by
doping the catalyst with In2 O3 [233c]. This was conrmed by a high
yield of ethanol during the HPR of acetic acid conducted at 2.1 MPa
total pressure between 220 and 380 C. Apparently, the involvement of InNi2 species, which was detected in the catalyst was the
main reason for the high selectivity for alcohol formation.
The Fe/SiO2 and Co/Kieselguhr solids were compared using
guaiacol as reactant [234]. The experiments were conducted in a
continuous xed bed reactor between 623 and 723 K and 1/WHSV
of 0.38 h in the ow of Ar + 90% H2 . The Co/kieselguhr catalyst gave
almost 100% conversion compared with about 77% for the Fe/SiO2
catalyst. Very high yields of methane over the former catalyst were
noted. At the same time, predominantly aromatic products were
formed.

6.1.2.4. Metal borides. Wang et al. [214218] prepared a series of


the amorphous (unsupported and unsulded) Co(Ni)Mo(W)B catalysts by chemical reduction of the corresponding metal salts with
sodium borohydride. The catalysts were used for the HDO (498 or
523 K, 4 MPa) of several model compounds found in biofeeds. During the HDO of phenol, the NiWB amorphous catalyst exhibited a
high HYD activity, giving cyclohexanol as the main product. In the
overall HDO, the dehydration of cyclohexanol to cyclohexene was
the rate-limiting step. It was suggested that amorphous Ni facilitated HYD activity, whereas WO3 acted as a Bronsted acidic site.
To certain amount, the addition of La improved catalyst activity by
enhancing the dehydration step. However, La in an excess blocked
active sites and decreased the catalyst surface area. Because of a
high HYD activity, no benzene formation was observed over the
La-doped NiWB catalysts. A similar effect of La was also observed
for NiMoB catalysts. Moreover, the activity of NiMoB catalysts was
further improved by the addition of Co [215]. However, the most
active NiMoB catalyst was prepared by the ultrasound assisted
reduction [217]. This was attributed to the higher concentration of
Mo4+ and unsaturated Ni active sites. Further increase in the activity of NiMoB catalysts was achieved by the addition of an optimal
amount of Co. Otherwise, Co in excess decreased activity [218]. Over
CoB, the conversion of phenol approached 100% with a selectivity of
91.4% to cyclohexanol [216]. The addition of Mo to CoB resulted in
a signicant increase in selectivity to cyclohexane. At 523 K, phenol
was converted mostly to cyclohexane and cyclohexanol, whereas
at 548, the main product was cyclohexane. In addition to phenol,
benzaldehyde and acetophenone were used as model compounds
over CoMoB catalyst under otherwise identical conditions [216].

6.2. Catalysts for HPR of vegetable oils biofeeds

6.1.2.5. Other catalysts. A series of Ni containing catalysts promoted with Cu were combined with various oxides (e.g., -Al2 O3 ,

Under thermal HCR conditions (no catalyst), vegetable oils are


converted to the mixtures of parafns, cycloparafns and aromatic hydrocarbons as well as relatively large amount of free
fatty acids [178180]. The presence of catalyst is necessary to
convert triglycerides in vegetable oils to hydrocarbons in boiling
range of diesel. For HPR applications, wide range of catalyst, both
conventional and unconventional formulations, have been tested.
High yields of green diesel from vegetable oils require catalysts
exhibiting high activity for HYD, HDO, HIS, HCR and decarboxylation/decarbonylation under typical HPR conditions. Conventional
HPR catalysts are suitable because they give high yields of diesel
fractions. However, cold ow properties (e.g., pour point, freezing
point, etc.) of the diesel may not comply with the specications of
commercial diesel. These properties may be improved by either
an additional treatment using a HIS catalyst or by using multifunctional catalysts. These requirements have been addressed
during the catalyst development as indicated by the number of
bifunctional catalysts shown in the summary of studies in Table 18.
6.2.1. Conventional HPR catalysts
Almost complete conversion of glycerides and corresponding
fatty acids during the HPR of sunower oil (360420 C, 18 MPa,
sulded commercial NiMo/Al2 O3 , continuous xed bed reactor),
was achieved [65]. The products distribution was strongly inuenced by temperature. For example, the concentration of n-alkanes
such as C17 and C18 , decreased from 53 to 5 wt.% by temperature
increase from 360 C to 420 C, while the concentration of aromatics
increased from a negligible amount to more than 5 wt.%. The concentration of i-alkanes increased by temperature increase as well.
Properties of the product obtained at 420 C approached those of

E. Furimsky / Catalysis Today 217 (2013) 1356

43

Table 18
Catalysts for HPR of biofeeds derived from vegetable oils.
Conventional catalysts

Feed

Conditions

Ref.

1. Sulded NiMo/Al2 O3
2. Sulded NiMo/Al2 O3
3. Oxidic and sulded NiMo/Al2 O3
4. CoMo/Al2 O3
5. Sulded NiMo/Al2 O3 and CoMo/Al2 O3
6. Sulded CoMo/Al2 O3
7. Reduced CoMo/Al2 O3
8. Sulded NiMo/meso-Al2 O3
9. Sulded NiMo, Ni and Mo all supported on
-Al2 O3
10. Sulded CoMo supported on organized
Al2 O3 and Al2 O3 , sulded CoMo/MCM-41
11. Sulded CoMo/MCM-41
12. Sulded CoMo, NiMo and NiW supported
on Al2 O3 and B2 O3 -Al2 O3
13. Sulded NiMo supported on SiO2 ,
SiO2 -Al2 O3 and different zeolites
14. Sulded CoMo supported on SBA-15,
SBA-16, DMS-1 and HMS and NiMo/Al2 O3
16. NiMo and CoMo supported on mesoporous
titano-silicates

Sunower oil
Rape seed oil
Rapeseed oil
Cotton seed oil
Waste cooking oil
Rapeseed oil
Sunower oil
Rapeseed oil
Triglycerides

Cont.; 360420 C, 18 MPa


Cont.; 360 C; 7 MPa
Batch; 350400 C; 120 MPa
Cont.; 305345 C; 3 MPa
Cont.; Cont.; 330390 C
Cont.; 310 C; 3 MPa
Cont.; 380 C; 46 MPa
Cont.; 260280 C; 3.5 MPa
Cont.; 310 C; 7 MPa

[65]
[15,16]
[235,236]
[203]
[237,240]
[135]
[241]
[242]
[243,244]

Triglycerides

Cont.; 310 C; 7 MPa

[177]

Triglycerides
Waste cooking oil and trapped grease

Cont.; 300 and 320 C; 212 MPa


Cont.; 300 and 350 C;

[245]
[246]

Jatropha oil

Cont.; 350 C; 4 MPa

[247,248]

Olive oil

Cont.; 250 C; 3 MPa

[249a]

Jatropha biofeed

Cont.; 300 and 360 C; 8 MPa

[205b]

Jatropha biofeed
Soyabean oil

Cont.; 330390 C; 3 Mpa


Batch; 400 C; 2.5 MPa

[249b]
[250]

Oleic acid and tripalmitin


stearic acid
Stearic acid

Batch; 325 C; 2.0 MPa


Batch; 400 C; 1 MPa
Batch; 300 C; 0.6 MPa

[179]
[254256]
[257,258a]

Stearic acid and tristearin


Rapeseed oil
Palmitic acids
Jatropha oil
Sunower oil
Sunower oil
Triglycerides

Batch; 300 and 360 C; H2 + N2


Batch; 300400 C; 5 and 11 MPa
Batch; 250 and 300 C; 1 and 4 MPa
543 and 573 K
250 C; 2 MPa
Cont.; 360450 C; 4.5 MPa

[258b]
[259a]
[259b]
[248]
[260]
[261a]
[261b]

Sunower oil
Methyl palmitate

Cont.; 2.1 MPa


Batch; 220 C; 2 MPa

[261c]
[261d]

Non-conventional catalysts
1. PtPd/Al2 O3 and NiMoP/Al2 O3
2. Unsulf. Ni/SiO2 -Al2 O3 , Pd/Al2 O3 , Pt/Al2 O3 ,
Ru/Al2 O3 , sulf. NiMo/Al2 O3 and CoMo/Al2 O3
3. Pt(5 wt.%)/Al2 O3 , Pd/Al2 O3 and Ni/Al2 O3
4. Pd/C and Pd/SBA-15
5. Ni, Ru, Pd, Pt, Ir, Os and Rh supported on
Al2 O3 , carbon, MgO, SiO2 and Cr2 O3
6. Ni/C and Pd/C
7. Pt/H-Y, Pt/H-ZSM-5 and sulded NiMo/Al2 O3
8. Pt/-zeolite oleic acid
9. Pt/H-ZSM5, RePt/H-ZSM and Pt/USY
10. Pd/SAPO-31
11. Pd/SBA-15 and Pd/HZSM-5
12. Ni-Mo carbide and nitride supported on
HSZM-5
13. Raney Ni, Ni/Al2 O3 and NiMo/Al2 O3
14. Ni supported on SiO2 , -Al2 O3 , SAPO-11,
HZSM-5 and HY.

petroleum-derived diesel fuel. This product had also excellent cold


ow properties.
Djega-Mariadassou et al. [235,236] used the oxidic and sulded
NiMo/Al2 O3 catalyst to study the effect of temperature and pressure on the HPR of biofeeds derived from maracuja, buritim tucha
and babassu sources. Rapeseed oil was studied over three different
NiMo/Al2 O3 catalysts by Simacek et al. [15,16], while Sebos et al.
[203] used cotton seed oil over commercial CoMo/Al2 O3 catalyst in
a trickle bed reactor. Waste cooking oil (WCO) was also investigated
as biofeed for the production of diesel over commercial NiMo/Al2 O3
catalyst [237239].
The results from the HPR of a waste cook oil over three commercial HPR catalysts are shown in Table 19 [240]. Catalysts A, B and C
used for HPR of waste cook oil included conventional NiMo/Al2 O3
for HDS and HDN, medium severity CoMo designed for the HYD and
mild HCR and NiMo HCR catalyst used for maximizing the yield of
middle distillates, respectively. For the experiments, biofeed was
spiked with DMDS to ensure a steady performance of catalysts.
Signicant increase in the H/C ratio during HPR conrmed an extensive saturation of double bonds. Thus, the waste cook oil contained
about 34 and 55% of C18:1 and C18:2 molecules, respectively. The
boiling point results indicated a higher HCR activity of catalyst A
compared with catalysts B and C. The H/C ratio suggests that catalyst A had also the highest HYD activity.
Kubicka and Horcek [135] used several rapeseed oils as HPR
feeds over the sulded CoMo/Al2 O3 catalyst. The feeds differed
in concentration of inorganic impurities, water, free fatty acids

and phospholipids. The experiments were carried out in xed bed


reactor (310 C, WHSV = 2 h1 , H2 pressure 3.5 MPa). Fig. 26 [135]
shows that the rened rapeseed oils were converted more efciently than the other feeds. An enhanced catalyst deactivation was
observed for the feeds with the high concentration of phospholipids
and alkalis. The continuous addition of the sulfur donating agent
Table 19
Quality comparison of products from HPR of waste cook oil (WCO) over A, B and C.
WCO

WCOa

Density (kg/L)
Sulfur (ppm)
Nitrogen (ppm)
Hydrogen (wt.%)
Carbon (wt.%)
H/C
Oxygen (wt.%)

0.8931
0.5
21.8
11.6
76.6
1.82
11.9

0.8955
27,200
220
11.4
81.9
1.67
3.9

0.7679
72
<1
14.8
85.0
2.09
0.2

0.8544
4580
215
12.7
81.3
1.87
5.4

0.7758
157
1
14.6
84.8
2.07
<1

BP dist., C
IBP (%)
5
10
50
90
95
FBP

451
576
598
609
613
615
634

99
526
590
607
610
610
627

49
213
271
305
419
462
536

104
211
271
366
537
574
621

101
196
271
307
460
475
571

Ref. [240].
a
WCO containing DMDS and an amine.

Catalyst

44

E. Furimsky / Catalysis Today 217 (2013) 1356

Fig. 27. Effect of Al2 O3 structure on performance of NiMo catalysts during conversion of rapeseed oil.
Ref. [242].

Fig. 26. Effect of purity of rapeseed oil (RO) on deoxygenation and yield of hydrocarbons; food grade RO,  degummed RO,  RO before degumming.
Ref. [135].

such as DMDS resulted in a signicant decrease in the deactivation


rate.
The CoMo/Al2 O3 catalyst was activated by reduction and used
for the HPR of sunower oil in a continuous system at 380 C,
46 MPa and LHSV of 1.0 h1 [241]. Under these conditions,
almost complete conversion of triglycerides was achieved via
HDO, decarboxylation and decarbonylation. Apparently, the long
term performance of either oxidic or prereduced form of conventional catalysts during the HPR of biofeeds still needs to be
demonstrated.
A mesoporous alumina (meso-Al2 O3 ) was compared with the
conventional Al2 O3 as the supports of sulded (at 340 C with
DMDS) NiMo catalysts [242]. The surface area of the supports was
255 and 520 m2 /g, respectively. The catalysts were used for the
HPR of rapeseed oil at 260280 C and 3.5 MPa total pressure. The
NiMo/meso-Al2 O3 exhibited much better performance than conventional NiMo/Al2 O3 catalyst (Fig. 27). Both catalysts contained
the same amount of Ni and Mo metals and were prepared by
the same procedure. The distribution of the Ni and Mo metals on
catalyst surface was similar as well. The signicantly higher specic
surface area and the total pore volume of NiMo/meso-Al2 O3 compared with the NiMo/Al2 O3 were the main contributors to much
higher activity of the former catalyst.
Kubicka and Kaluza [243] compared sulded NiMo, Ni and
Mo catalysts (all supported on -Al2 O3 ) during the conversion
of triglycerides to hydrocarbons at 310 C and 7 MPa. For the
overall conversion, the activity decreased in the following order:
NiMo/Al2 O3 > Mo/Al2 O3 > Ni/Al2 O3 . The NiMo catalysts were the
most selective to hydrocarbons via both HDO and decarboxylation, while decarboxylation and HDO were dominant reactions
over the Ni/Al2 O3 and Mo/Al2 O3 , respectively. It was further conrmed that the selectivity for hydrocarbons could be inuenced by
the change in distribution of Ni species in the NiMo/Al2 O3 catalysts [244]. In this regard, the surface structure may be ne tuned
during the catalyst preparation. In addition, Kubicka et al. [177]

used sulded CoMo catalyst supported on an organized aluminasupport (OAS). The CoMo/OAS catalysts exhibited higher activity
than CoMo/MCM-41 and CoMo/Al2 O3 catalysts. The overall conversion involved two main routes, i.e., HDO and hydrodecarboxylation.
The relative contribution of these pathways was inuenced by temperature, pressure and type of the support.
Rapeseed oil was used as the feed to study activity of the CoMo
catalysts supported on a series of mesoporous molecular sieves
(MCM-41) with varying Si/Al ratio [245]. The experiments were
conducted in a ow reactor at 300 and 320 C, H2 pressure of
212 MPa and LHSV of 14 h1 . The focus was on the conversion
of triglycerides. The activity of catalysts supported on the unmodied MCM-41 support was rather low. Incorporation of Al into the
framework of MCM-41 increased both conversion and selectivity
to n-heptadecane and n-octadecane. However, the conversion of
triglycerides was lower and selectivity to oxygenates higher than
those achieved over CoMo/Al2 O3 . Benecial effect of incorporation
of silica into MCM-41 support was even lower than that of alumina.
Moreover, the stability of MCM-41 in the presence of water may be
affected.
The HDO of waste cooking oil and trapped grease was conducted
over the conventional sulded CoMo, NiMo and NiW catalysts in
a continuous xed bed reactor [246]. The NiMo and NiW catalysts
exhibited a high and steady activity, whereas the CoMo catalyst was
prone to deactivation. Better catalyst performance was observed
at 350 C than at 300 C. The HIS reactions occurred over catalysts
supported on an acidic support such as B2 O3 Al2 O3 . Because the
tests were conducted without a sulfur donating agent being present
in the feed, a small amount of sulfur from catalyst dissolved in the
product streams.
Several NiMo catalysts supported on different supports were
compared during the HPR of the biofeed derived from Jatropa oil
[247] the composition of which is shown in Table 9 [14]. Experiments were performed in a continuous xed-bed reactor at 350 C
under 4 MPa of H2 . The properties of the liquid hydrocarbon products are shown in Table 20 [247]. Triglycerides and free fatty acids
were converted to mixed parafns, however different amounts
of free fatty acids, were still present. In the case of NiMo/SiO2
catalyst, the n-C18 H38 , n-C17 H36 , n-C16 H34 , and n-C15 H32 were
the predominant products. These hydrocarbons had high melting
points. Consequently, the product had a high pour point. When
SiO2 Al2 O3 was used as the support for the NiMo catalyst, the
pour point of the liquid product decreased to 10 C from about
20 C. This was the result of an enhanced conversion of C15 C18

E. Furimsky / Catalysis Today 217 (2013) 1356

45

Table 20
Composition (wt.%) and property of liquid hydrocarbons from HPR Jatropha oil (350 C; H2 of 4 MPa; H2 /oil of 800 mL/mL; LHSV, 7.6 h1 ).
Catalyst

Composition

NiMo/SiO2
NiMo/Al2 O3
NiMo/SiO2 .Al2 O3
NiMo/H-Y
NiMo/H-ZSM-5
Commer. diesel

Property

C5 C10

C11 C20

C20 +

iso/n

Pour point ( C)

Densitya (g/mL)

Viscosityb (MPa/s)

0.8
2.5
9.8
48.9
77.8
8.2

99.1
97.2
89.6
50.9
22.1
88.1

0.1
0.3
0.6
0.2
0.1
3.7

0.03
0.08
0.26
0.87
1.21
0.28

20
10
10

0.79
0.79
0.78
0.76
0.75
0.82

8.28
7.41
4.13
2.08
1.16
3.69

c
c

15

Ref. [247].
a
At 25 C.
b
At 30 C.
c
Lower than 25 C.

n-parafns to i-parafns because SiO2 Al2 O3 had a proper acidic


strength. Thus, the i-parafns/n-parafns ratio approached 0.26.
Also the viscosity of the liquid hydrocarbons produced over the
NiMo/SiO2 Al2 O3 catalyst was quite similar as that of the commercial diesel fuel, although a difference in the density was noted.
Nevertheless, the liquid products obtained over NiMo/SiO2 Al2 O3
may be used directly in the current diesel engines. It is, therefore
evident that the NiMo/SiO2 Al2 O3 catalyst was the most suitable
when the production of diesel fuel was the objective. However,
for high yields of biogasoline, NiMo/H-Y and NiMo/H-ZSM-5 would
be the catalysts of the choice. The most suitable NiMo/SiO2 Al2 O3
catalyst was used to compare the yields and properties of the
hydrocarbon fractions from Jatropha biofeed with those obtained
from canola and palm oil biofeeds. In this case, the yields of biohydrogenated diesel (BHD), bio-LPG fuel (C3 H8 + C4 H10 ) and fuel
gas (CH4 + C2 H4 ) were determined. As the results in Table 21 [248a]
show, large differences in product composition among the biofeeds
were not observed.
The NiW/ZSM-5 catalyst with ZSM-5 support possessing hierarchical structure and intra-crystalline mesoporosity was used for
the HCR of triglycerides and free fatty acids obtained from algae
and Jatropha seeds [248b]. The yield of kerosine fraction (C9 C15 )
ranged from 40 to 45% with the iso-/normal (i/n) parafns ration
between 2 and 6. The NiMo catalyst supported on the same support gave 4050% kerosene with i/n of 313. For the NiMo catalyst
supported on high surface area semi-crystalline ZSM-5, the yield of
kerosine approached 77% with the i/n ratio of 2.5.
The sulded CoMo catalysts supported on different mesoporous
silicates (e.g., SBA-15, SBA-16, DMS-1, and HMS) were compared
with conventional NiMo/Al2 O3 catalyst during the HPR olive oil
(xed-bed reactor, 250 C, 3 MPa) [249a]. All CoMo catalysts were
much more active than the NiMo/Al2 O3 catalyst. Considering the
yield of desirable products, the CoMo/SBA-16 catalyst was the most
active, while the CoMo/HMS was the least suitable. Also, the former
catalyst exhibited a high activity for both HDO and HDAr.
Mesoporous titanosilicate (MTS) prepared by solgel method
and -Al2 O3 were used as supports for sulded CoMo and NiMo
catalysts (jatropha biofeed, 300 and 360 C, 8 MPa, LHSV of 2, 4 and

8 h1 ) [205b]. At 360 C, the overall conversion of triglycerides was


85 and 66% over the NiMo and CoMo catalysts supported on MTS,
respectively. For these catalysts, almost complete triglycerides conversion was achieved at 360 C. Compared with -Al2 O3 supported
catalysts, the MTS supported catalyst gave higher yield of kerosene
distillate (C9 C14 ) and much lower yield of heavy products (>C18 )
because of a favorable surface acidity. Moreover, the fuel quality of
the products obtained over the MTS supported catalysts was much
better as indicated by the ve times greater iso-/n-parafn ratio
(Table 22) [205b] compared with the -Al2 O3 supported catalysts.
6.2.2. Non-conventional catalysts
Bifunctional catalysts consisting of noble metal catalysts and
supports varying widely in surface acidity as well as metal carbides, nitrides and phosphides have been receiving attention. It was
evident that the bifunctional catalysts were of the primary interest. In most studies, conventional HPR catalysts were also used for
comparison.
6.2.2.1. Noble metals containing catalysts. The studies in which nonconventional catalysts are compared with conventional catalysts
under identical HPR conditions are of a particular interest. Gong
et al. [249b] reported results on the HPR of Jatropha biofeed over
PtPd/Al2 O3 and NiMoP/Al2 O3 catalysts (continuous xed bed reactor, 330390 C, 3 MPa of H2 ). The yields of hydrocarbons and their
properties are shown in Table 23. Liquid products had low acid
number, suitable density and viscosity and unusually high cetane
index. These liquids could be used as blending components with
diesel of petroleum origin, although only a limited amount could
be added because of a high cloud point. The mass balance of hydrocarbons in products showed that over PtPd/Al2 O3 , predominant
route involved decarboxylation and/or decarbonylation, whereas
HDO was the main route over NiMoP/Al2 O3 catalysts. For the
NiMoP/Al2 O3 catalyst, slow deactivation resulted from a gradual
conversion of the suldic form of metals to oxidic form.
Six different catalysts investigated by Veriansyah et al. [250]
comprised conventional sulded NiMo/Al2 O3 and CoMo/Al2 O3
catalysts as well as the unsulded/prereduced Ni/SiO2 Al2 O3 ,

Table 21
Yield of products and properties of BHD (NiMo/SiO2 Al2 O3 ; 350 C; H2 of 4 MPa).
Biofeed

Jatropha
Canola
Palm
Ref. [248].
a
C3 H8 + C4 H10 .
b
CH4 + C2 H6 .
c
At 25 C
d
At 30 C

Yield (wt.%)

BHD property

BHD

LPGa

Fuel gasb

COx

H2 O

Iso/n

Pour point ( C)

Densityc (g/mL)

Viscosityd (mPa/s)

83.5
81.4
82.1

4.9
5.7
5.4

0.2
0.3
0.3

2.4
2.1
2.3

8.9
9.5
9.3

0.26
0.25
0.27

10
10
15

0.78
0.79
0.78

4.13
4.22
3.82

46

E. Furimsky / Catalysis Today 217 (2013) 1356

Table 22
Distribution of products (vol.%) from HPR of jatropha biofeed.a
Support

Al2 O3
MTS

Conversion (%)

Yield of products (vol.%)

98.7
99.9

<C8

C9 C14

>C18

C15 C18

C17 /C18

i/nb

3.8
1.2

10.5
18.1

10.4
4.3

75.3
78.3

0.65
0.25

0.3
1.5

Ref. [205b].
MTS, mesoporous titanosilicate.
a
Sulded CoMo supported on -Al2 O3 and MTS, 360 C, 8 MPa, 2 h1 .
b
i/n, ratio of i-parafns/n-parafns.

Pd/Al2 O3 , Pt/Al2 O3 and Ru/Al2 O3 catalysts. The performance of


these catalysts was compared during the HPR of soyabean oil in
batch reactor at 400 C and 2.5 MPa. The C18 species accounted for
more than 80 wt.%, while C16 for about 11 wt.%. The following catalyst activity order was established: NiMo/Al2 O3 (92.9%) > Pd/Al2 O3
(91.9%) > CoMo/Al2 O3 (78.9%) > Ni/SiO2 Al2 O3 (60.8%) > Pt/Al2 O3
(50.8%) > Ru/Al2 O3 (39.7%) at the catalyst/feed ratio of 0.044, as
well as Ni/SiO2 Al2 O3 (95.9%) > NiMo (91.9%) > Pd (90.9%) > CoMo
(79.9%) at the ratio of 0.088. For two most active catalysts, high
diesel selectivity (e.g., 9498%) was observed, while naphtha selectivity was only 37% (at the catalyst/feed ratio of 0.044). This
suggests that HCR reaction were negligible. The product obtained
over NiMo/Al2 O3 catalyst had a high content of i-alkanes compared
with that obtained over Pd/Al2 O3 . For the CoMo/Al2 O3 catalyst, the
naphtha and diesel selectivity approached 17.7 and 82.3%, respectively indicating an involvement of HCR reactions. For less active
Ni, Pt, Ru catalysts, oxygen content in products varied from 4.7 to
9.3 wt.%. Propane, one of the main products from HPR of glycerides,
was produced in a larger quantity over the Pt catalyst (7.8 mol%).
Madsen et al. [179] compared three catalysts, i.e., Pt/Al2 O3 ,
Pd/Al2 O3 and Ni/Al2 O3 during the HDO of the mixture of oleic
acid and tripalmitin (molar ratio of 1/3) at 325 C, 2.0 MPa of H2 in
an autoclave. Under these conditions, decarboxylation and decarbonylation of the ester and carboxylic acid, yielding odd carbon
numbers hydrocarbons, were highly favored. The Pd/Al2 O3 catalyst
Table 23
Yields and properties of liquids obtained over PtPd/Al2 O3 and NiMoP/Al2 O3 (3 MPa
H2 ; LHSV of 2 h1 ; H2 /feed of 600).
Yatropha feed

Products
PtPd/Al2 O3

Yield
Liquid hydrocarbonsa
Gaseous hydrocarbons
Water
Property
Density, kg/L at 15 C
Viscosity, mPa s at 25 C
Acid number, mg KOH/g
Cloud point, C
Cetane index
Boiling range, C
IBP
50%
FBP
Composition
C14
C15
C16
C17
C18
C18+

81.2
5.6
4.3

0.91 (0.770.80)b 0.78


21.17 (24.5)
3.21
8.48
<0.1
4
20
45 (5170)
93
271
294
484

15.6

84.2
0.4

<0.4
13.9
<0.5
81.4
1.5
2,1

NiMoP/Al2 O3
83.9
5.6
7.5
0.78
3.25
<0.1
18
97
120
304
475
<2
7.6
7.3
37.3
36.5
2.4

Ref. [249b].
a
Iso-parafns accounted for <0.3 and 7% in liquids obtained over PtPd/Al2 O3
and NiMoP/Al2 O3 , respectively.
b
Results in brackets are those of a commercial diesel fuel.

was slightly more active than Pt/Al2 O3 catalyst, while the activity
of Ni/Al2 O3 was much lower.
The Pd/C based catalysts have been tested for HPR of various biofeeds. For example, Arend et al. [251] reported that conversion of
the pure oleic acid (no solvent) and selectivity to heptadecane and
heptadecenes over a granular Pd(2 wt.%)/C catalyst was favored at
low reaction temperatures, a high H2 ow and a low feed/catalyst
ratio. The structure and performance of the Pd/C catalysts during the HPR of free fatty acids was studied extensively by Murzin
et al. [252258]. The decarboxylation of four pure fatty acids (e.g.,
C17 C20 ) over 1 wt.% Pd/C (semibatch reactor, 360 C; 1 MPa of Ar or
5% H2 + Ar) proceeded at similar rates, however an extensive catalyst deactivation was observed for the C22 fatty acid [229,253]. This
was attributed to a low purity of the acid caused by the presence of
phosphorus. The Pd supported on mesoporous carbon exhibited a
steady performance in the xed bed down ow reactor using a pure
stearic acid as the feed [254,255]. The main liquid-phase product
was heptadecane, while the main gaseous products were CO and
CO2 . Without H2 , the catalyst deactivation due to coke formation
was quite evident. Under the same experimental conditions, similar deoxygenation trends were observed over Pd/SBA-15 catalysts
[256].
Snare et al. [257] evaluated 20 catalysts containing different
metals (e.g., Ni, Ru, Pd, Pt, Ir, Os and Rh) and supports (e.g., Al2 O3 ,
carbon, MgO, SiO2 and Cr2 O3 ) during the conversion of stearic
acid in a semibatch reactor at 300 C and 6 bar. Apparently, little of gaseous H2 was present, however, under the conditions
employed, a hydrogen transfer from the solvent (dodecane) to
reactant molecules almost certainly took place. Otherwise, the
conversion was enhanced in the presence of gaseous H2 (H2 + He
mixture) [258a]. The extensive characterization of the catalysts was
part of the study as well. For the most active Pd(5%)/C catalyst,
the conversion approached 100% with C17 selectivity almost 99%.
Heptadecane accounted for 95% of C17 mixture. For the Pt(5%)/C
catalyst, the conversion was 86% with C17 selectivity of 95%. Over
bimetallic Pd(8%)Pt(2%)/C, the overall conversion of about 60% with
C17 selectivity of 96% was observed. About 33 and 48% conversions
were achieved for the Pd/C catalysts containing either 1 or 10%
of the metal, respectively. For the most active catalysts, the HDO
of stearic acid to C17 products (mostly n-C17 ) was far predominant reaction. Thus, only traces of C18 hydrocarbons were detected.
In comparison with these catalysts, activity of the conventional
NiMo/Al2 O3 was low, i.e., less than 9%. The Ru(5%)/MgO catalyst was
also very active (96% conversion) with 99% selectivity to symmetric C35 ketone. The ketonization of stearic acid was also catalyzed
by Ir/Al2 O3 , Pd/Al2 O3 , Pt/Al2 O3 , Ru/SiO2 and Rh/SiO2 .
The conversion of stearic acid and tristearin was studied in a
semi-batch autoclave over 20 wt.% Ni/C and 5 wt.% Pd/C catalysts in
the ow 10% H2 + N2 at 300 C and 1.5 h as well as at 360 C and 6 h
for stearic acid and tristearin, respectively [258b]. For the experiments, the reactants (1.75 g) were dissolved in dodecane (25 g).
Because of a higher acidity and higher temperature used with the
Ni/C catalyst, a high yield of hydrocarbons in C10 C17 range were

E. Furimsky / Catalysis Today 217 (2013) 1356

47

Fig. 29. Effect of Si/Al ratio on C17 and C18 hydrocarbons selectivity.
Ref. [261a].
Fig. 28. Effect of zeolite type on yield of hydrocarbons during conversion of rapeseed
oil.
Ref. [259a].

produced, while Pd/C catalyst exhibited a high selectivity to C17 .


The former catalyst was more prone to deactivation by coke.
Bifunctional Pt/H-Y and Pt/H-ZSM-5 catalysts were compared
with sulded NiMo/Al2 O3 (batch reactor, 300400 C, 511 MPa,
3 h) during the HPR of rapeseed oil [259a]. Fig. 28 shows the
yields of products over the Pt-zeolite catalysts. The less acidic
Pt-HY gave higher yields of diesel (C13 C22 ) compared with
gasoline (C5 C12 ) fractions. However, the highest yield of diesel
hydrocarbons (mostly straight chain) were obtained over the
NiMo/Al2 O3 catalyst. This fraction had the highest cetane number
but its cold ow properties were rather poor compared with the
diesel fractions obtained over bifunctional Pt-zeolite catalysts.
Reaume and Ellis [259b] used the -zeolite with Si/Al molar
ratio of 25 and surface area of 680 m2 /g for isomerization of oleic
and palmitic acids while the Pt(0.5)/-zeolite for HIS of these acids.
Tests were performed in a batch reactor at 250 C and 1 MPa H2 as
well as at 300 C and 4 MPa H2 , respectively. The isomerization had
no positive effect on cloud point reduction. At the same time, HIS
had an adverse effect for oleic acid and a positive effect for palmitic
acid. With respect to the cloud point reduction of the corresponding
FAMEs, the optimal combination was the mixture of oleic acid and
palmitic acid, (e.g., 55/45 ratio). At the optimum, the cloud point
reduction of 7.5 C was achieved.
Murata et al. [248] compared the Pt/H-ZSM5 and Pt/USY catalysts during the HPR of jatropha oil (Table 9) with the aim to produce
C15 C18 hydrocarbons directly. Over Pt/USY, almost 100% conversion of the biofeed, with 90% selectivity to C10 C20 alkanes, was
achieved at 573 K. Other major products included C1 C10 hydrocarbons and CO2 . The highest conversion was achieved at 543 K
over Pt/H-ZSM-5 with C15 C18 alkane selectivity being 79%. However, at high jat/cat ratios, i.e., 2 and 10, the C15 C18 alkane yields
fell to 13.5 and 2.7%, respectively. Rhenium-modied Pt/H-ZSM-5
catalysts were found to be much more effective. Thus, even at the
jat/cat ratio of 10, and 80% conversion, 70% C18 selectivity were
achieved.
The bifunctional nature of Pd/SAPO-31 catalyst was demonstrated during experiment carried out in a laboratory ow reactor
using sunower oil as the feed [260]. Products contained mostly
n-alkanes and i-alkanes of C17 and C18 carbon number. The catalyst exhibited a high initial conversion and HIS selectivity but its
deactivation was evident after several hours on stream.

Fig. 29 [261a] shows the effect of the Si/Al ratio on the yield of C17
and C18 hydrocarbons during the HPR of sunower oil over Pd/SBA15 catalysts (250 C and 2 MPa of H2 ). An addition of small amounts
of Al had dramatic effect on the products yield and distribution.
Thus, the yield of C18 hydrocarbons increased with the decreasing
Si/Al ratio. At the same time, the yield of C17 increased and reached
maximum at about Si/Al ratio of 300 before it decreased with
further Si/Al ratio decrease. The Pd/HZSM-5 (22) catalyst was compared with the Pd/SBA-15(22) catalyst. At 300 C, aromatization
was the dominant reactions over the former catalyst.
6.2.2.2. Metal carbides and nitrides. Bimetallic (Ni/Mo) carbide or
nitride supported on ZSM-5 were prepared by Wang et al. [261b].
The method involved the impregnation of ZSM-5 by incipient wetness using aqueous solutions of Ni(NO3 )2 and (NH4 )6 Mo7 O24 4H2 O
salts. The obtained solid was dried for 12 h at 120 C before calcination at 400 C for 6 h. The carburization to obtain carbides was
carried out in a ow of CH4 /H2 at 730 C, while nitridation to obtain
nitrides in the ow of ammonia at 700 C. The catalysts were
used for the HPR triglycerides (4.5 MPa and 360450 C, continuous xed bed reactor). The product obtained at low temperature
(360 C) contained predominantly carboxylic acids with hydrocarbons accounting for less than 5% of the total product. The yield
of hydrocarbons (gasoline and diesel) increased with increasing
temperature and decreasing LHSV. For both the Ni/Mo carbides
and nitrides, the presence of H2 O, CO and CO2 in the product
conrmed the involvement of HDO, decarbonylation and decarboxylation reactions.
6.2.2.3. Other catalysts. Onyestyk et al. [261c] used nonpyrophoric Raney-Ni (27 wt.% Ni) to study conversion of the
rened sunower oil (food grade) containing 94 wt.% C18 fatty acids
(mainly linoleic and oleic). The activity of this catalyst was compared with the commercial NiMo/Al2 O3 modied with P and Si
as well as with the Ni (27 wt.%)/Al2 O3 catalysts. The effect of Pb
and Sn metals on the activity of the Raney Ni catalyst was investigated as well. The tests were conducted in the continuous xed bed
reactor using different ratios of the feed/H2 at 2.1 MPa and WHSV
of 1 h1 , 2 h1 and 3 h1 . The NiMo/Al2 O3 showed a similar conversion as the Ni/Al2 O3 producing octadecane and heptadecane as
the main products, while over the latter catalyst decarbonylation
and decarboxylation followed by methanation of CO and CO2 , were
evident. The methanation reactions were also observed over the

48

E. Furimsky / Catalysis Today 217 (2013) 1356

67 wt.% gasoline fraction, 15% jet fuel and 15% diesel fuel. The gasoline and jet fuel fractions consisted of n-alkanes and i-alkanes.

7. Coprocessing of biofeeds

Fig. 30. Boiling range of (a) algae biofeed, (b) product of HPR of (a), and (c) conventional petroleum.
Ref. [262].

non-pyrophoric Raney-Ni. These reactions could be suppressed by


the addition of In2 O3 .
Zuo et al. [261d] prepared the sulfur free Ni catalysts supported
on SiO2 , -Al2 O3 , SAPO-11, HZSM-5 and HY by incipient wetness
impregnation. The catalysts were tested during the HDO methyl
palmitate. The C16 alkanes were formed via stepwise conversion
involving the HYD of hexadecanal to hexadecanol, followed by
deHYD/HYD. The decarbonylaton/decarboxylation of hexadecanoic
acid to C15 alkane occurred in parallel. The Ni (7 wt.%)/SAPO-11
catalyst with weak and medium acidity showed superior activity,
i.e., 93% yield of C15 + alkanes at 493 K and 2 MPa in a semibatch
reactor.
6.3. Catalysts for HPR of algae biofeeds
In spite of challenges presented by high content of nitrogen in
the biofeeds obtained from algae biomass, a limited information
on catalyst development for HPR is rather surprising in view of
high hopes given to this source of the future biofuels. Thus, adverse
effects of nitrogen compounds on HPR catalysis have been well documented [80,99]. A detailed account of various algae, as potential
source of biofuels were given in the review published by Tran et al.
[43a]. The elemental analyses of some biofeeds derived from the
algae sources are shown in Tables 68.
Early studies on the HPR of biofeed derived from botryococcus braunii algae were conducted by Hillen et al. [262264]. The
biofeed was extracted from the algae by acetone and puried
by column chromatography using Al2 O3 . This may be the reason
for rather low nitrogen content of the biofeed (0.17 wt.%) compared with the other biofeeds of the algae origin (Tables 68). The
experiments were conducted over sulded CoMo/Al2 O3 catalyst
under rather severe conditions, i.e., between 400 and 440 C and
20 MPa of H2 . The reactor consisted of a 6.5 m coiled stainless-steel
packed with 120 g of catalyst. The feeding parameters indicate a
low feed/catalyst ratio and a long contact time. These conditions
ensured a high conversion of the biofeed. This is indicated by boiling point ranges of the biofeed and product in comparison with a
petroleum crude, shown in Fig. 30 [262]. The product comprised

The co-processing of biofeeds with petroleum fractions of a


similar boiling range under HPR conditions has been receiving
attention. In petroleum reneries, potential of both reforming and
FCC processes for coprocessing with biofeeds has been explored
with some positive results [265a,b]. For these processes, the
biofeeds of vegetable origin are the most suitable. Because the focus
of the review is on HPR, little attention is paid to the coprocessing
via alternative routes.
Pinheiro et al. [266a] used model oxygenates such as 2-propanol,
cyclopentanone, anisole, guaiacol, propanoic acid and ethyldecanoate as additives (0.5 wt.% oxygen) to SRGO of petroleum origin
to study their effect on HDS and HDN. Experiments were conducted
at 330 C and 5 MPa over sulded CoMo/Al2 O3 catalyst in a continuous system. Little inhibiting effect on HDS was observed for
the water producing reactants such as 2-propanol, cyclopentanone,
anisole, and guaiacol compared with propanoic acid and ethyldecanoate. Apparently, the inhibition was caused by the hydrogen
depletion due to methanation reaction involving CO and CO2 . Coprocessing of guaiacol with SRGO over CoMo/Al2 O3 catalyst below
320 C affected HDS rate however, above 320 C, complete HDO
of guaiacol was observed and HDS proceeded without inhibition
[266b].
The NiMo/Al2 O3 catalyst was tested using two mixtures of diesel
oil + rapeseed oil, i.e., containing 10 and 20 wt.% of rapeseed oil
[267a]. Under the conditions applied (320380 C, 35 MPa and
LHSV of 2 h1 ), it was possible to attain a high conversion of vegetable oil to diesel fuel hydrocarbons. Under similar conditions, the
mixture of straight run gas oil and rapeseed oil was investigated by
Aribert et al. [267b] over sulded CoMo/Al2 O3 catalyst. Apparently,
additional HPR may be necessary to meet the cold ow properties
specications of commercial diesel fuel.
Pretreatment of pyrolysis derived biofeed may improve efciency of the subsequent upgrading step as it was shown by Lappas
et al. [268]. In their study, quality of primary bio oil was enhanced
by thermal treatment without catalyst in the presence of H2 . In
this case, almost 85% oxygen removal from the bio oil could be
achieved with the nal liquid containing only 6.5 wt.% of oxygen.
Compared with Tables 3 and 8, this is rather signicant removal of
oxygen. Based on the non-treated feed, the yield of upgraded liquid
approached 42 wt.%. The gasoline, diesel and VGO fractions derived
from the upgraded bio-liquid were suitable for blending with the
corresponding petroleum liquids for further upgrading using HPR.
A potential blending component with petroleum feeds may also
be prepared by a mild HPR of pyrolysis bio oil, i.e., down to about
7 wt.% of oxygen [269]. However, when blending with the FCC feed
was considered, a decrease in the oxygen content down to about
20% was adequate for achieving a smooth HPR operation [270a].
Sebos et al. [203] studied the mixture of 10 wt.% of cottonseed oil in a petroleum diesel fraction over CoMo/Al2 O3 catalyst
(305345 C, 3.0 MPa and WHSV of 525 h1 ). The catalyst exhibited steady performance during 450 h on stream. Under these
conditions, almost 100% conversion of triglycerides in the mixture
was achieved. The cetane number of the nal product increased by
3 units compared with the original petroleum diesel.
Sulded NiMo/Al2 O3 catalyst was used by Huber et al. [168]
for the HPR of sunower oil, petroleum derived heavy vacuum oil
(HVO) and mixture of the sunower oil with the HVO (550 wt.%
of sunower oil) in a stainless steel tubular reactor (300450 C,
5.0 MPa and LHSV of 4.97 h1 ). The HVO contained 0.35 wt.% of sulfur and 1600 ppm of nitrogen. The yield of gases (mainly CO, CO2

E. Furimsky / Catalysis Today 217 (2013) 1356

and propane) increased with increasing temperature, similarly as


the yields of the 65150 and 150250 C fractions. The percentage
of straight chain alkanes in the 150250 C cut was less than 16%,
but increased with increasing sunower oil content. The yield of
the 250380 C fraction, containing mostly n-C15 n-C18 hydrocarbons, increased with increasing sunower oil content in the feed
mixture.
During co-HPR of the mixtures of waste cooking oil and
petroleum gas oil, the yield of diesel fraction (250380 C) over
NiMo/Al2 O3 catalyst varied between 85 and 95% compared with
less than 50% over the NiW/SiO2 Al2 O3 catalyst [270b]. For the latter catalyst, the lower yield of diesel fraction was offset by a high
yield of kerosine fraction (140250 C). Under conditions applied
(xed bed reactor; 340380 C; 5 MPa), HCR reactions were more
evident on more acidic NiW/SiO2 Al2 O3 catalyst. Under similar
conditions, the yield of C15 C18 hydrocarbons over of the sulded
NiW/SiO2 Al2 O3 , CoMo/Al2 O3 and NiMo/Al2 O3 catalysts 80.8, 49.2
and 97.9%, respectively [270c]. However, the iso- to n-parafns (i/n)
ratio was 2236 times higher for the NiW/Al2 O3 catalyst compared
with that for NiMo/Al2 O3 catalyst. Jatropha oil in mixture had little
effect on HDS of gas oil.
Three HPR catalysts were compared with respect to their selectivity (naphtha, kerosene/jet, diesel) during the co-HPR of the 70/30
mixture of petroleum VGO and vegetable oil [271]. The yields of four
fractions such as 360 C+, 180360, 170270 and IBP 200 C in the
products were compared with the contents of the same fractions
in the feed. The pre-HYD of VGO prior to HPR of the mixture gave
higher quality products.
A mild HPR of pyrolysis oil to moderate levels of oxygen may
be coupled with coprocessing with petroleum fractions [197]. In
this regard, the effect of HPR severity on quality of products was
investigated in batch reactor over a conventional HPR catalyst. The
results indicated that renery blendstocks meeting specications
of commercial fuels may be produced via integration of mild HDO
of biofeed with coHPR of petroleum feeds.
The coHPR of the mixture containing 5 wt.% of rapeseed oil with
a petroleum derived vacuum distillate was carried out (400 and
420 C and 18 MPa) over sulded NiMo/Al2 O3 catalyst [272a]. The
components indicating incomplete transformation of rapeseed oil
(e.g., free fatty acids and triglycerides) to hydrocarbons, were completely removed. The coHPR products obtained at 400 C had a
much higher content of n-C17 and n-C18 alkanes than those form the
HPR of the petroleum feed. At 420 C, the content of the n-alkanes
decreased while that of i-alkanes increased.
The blends of rened palm oil (5 and 10 wt.%) with HGO (0.5%
sulfur) were used by Vonortas et al. [272b] and Templis et al. [272c]
as feeds for HPR over CoMo/Al2 O3 catalyst (330365 C, 3.3 MPa
and WHSV of 0.85, 1.0, and 1.4 h1 ). Attempt was also made to
estimate hydrogen consumption. The addition of palm oil up to
5 wt.% abruptly decreased the rate of HDS, while further increase
to 10 wt.% had little incremental deactivating effect.
The blend of LCO and soybean oil in concentration from 0 to
30% was investigated by Yunqui [273a] over NiMoP/HUSY-Al2 O3
catalyst at 370 C and 4 MPa total pressure. The cetane number of
the produced diesel increased from 32.7 to 39.2 with the increasing
content of the soybean oil from 0% to 30%. However, both sulfur and
nitrogen removal were slightly affected.
The mixtures of heavy gas oil and waste cooking oil (90/10
and 70/30) were used for HPR (310350 C, 8 MPa and LHSV of
1.0 h1 ) over the sulded NiMo/Al2 O3 catalyst [273b]. The effect of
temperature on the rates of HDS, HDN, HDO and DDO as well as
pour point of liquid products and consumption of hydrogen, were
investigated. Heteroatoms removal increased with increasing temperature, while the content of waste oil had little effect. Waste oil
in the mixture benecially inuenced cold ow properties of liquid
products. At the same time, the hydrogen consumption increased.

49

8. Hydroprocessing of biofeeds in aqueous phase


The biofeeds produced during hydrothermal liquefaction are
formed in an aqueous medium. Separation of water soluble components from the aqueous phase is inefcient. A similar situation
may be encountered during the handling of biofeeds from biomass
pyrolysis. It may be advantageous to convert polar compounds in
biofeeds to hydrocarbons directly in the same environment [41].
If the HPR method is used, most of the polar components can be
converted to hydrocarbons. Then, the separation of hydrophobic
hydrocarbon phase from the aqueous phase is rather simple. The
temperatures employed during the aqueous HPR range from 200
to 450 C [20]. This temperature range covers both, sub-critical and
super critical water temperature regions. Thus, the supercritical
temperature of water is 374 C (647 K). These methods of the
biofeeds upgrading have been gradually attracting the attention
as evidenced by research studies appearing recently in the literature [274281a]. Obviously, rather different catalyst formulations
are needed for HPR to withstand potential adverse effects of aqueous conditions. Thus, for conventional catalysts, the transformation
of -Al2 O3 to boehmite leading to catalyst deactivation may not
be avoided. This problem may be alleviated by replacing -Al2 O3
support with more hydrophobic supports (e.g., carbons) [281b].
Besides HPR, there are other catalytic options for the aqueous phase
conversion of biomass to hydrocarbons and hydrogen [20].
8.1. Sub-critical conditions
Under sub-critical conditions, a high pressure is required to
ensure that most of the water in the system is in a liquid form. This
improves the interaction of water molecules with reactants. Further enhancement in conversion may be achieved in the presence
of H2 and catalyst.
The study published by Peng et al. [275] contributes to the
fundamental understanding of the HPR of oxygenates in an aqueous environment under sub-critical conditions. The reactants such
as 1-propanol, 2-propanol, 1,2-propanediol, 1,3-propanediol and
glycerol (10 wt.% in water) were studied in a batch reactor at
473 K, 4 MPa of H2 and 0.3 g of 3 wt.% Pt/Al2 O3 . Under these conditions, the direct cleavage of C C and C O bonds was not observed.
For 2-propanol and 1,2-propanol, deHYD to ketone was the main
reaction while for 1,3-propanol and glycerol, the C O bond was
cleaved by dehydration. For alcohols with the terminal hydroxyl
group, the C C bond cleavage occurred in steps via deHYD to
aldehyde followed by either disproportionation and subsequent
decarboxylation or decarbonylation. Reactivity of the alcohols
increased with increasing number of hydroxyl groups. Thus, the
following overall reactivity order was established: glycerol 1,3propanol > 1,2-propanol > 1-propanol 2-propanol.
The aqueous phase HDO of propanoic acid was used by Chen
et al. [276] to compare Ru/ZrO2 and RuMo/ZrO2 catalysts in a
trickle-bed reactor after the catalysts (6 g of 2040 mesh) were
activated in situ in the ow of H2 at 300 C for 3 h. In the temperature range of 150230 C, the total pressure of 6.4 MPa ensured
an aqueous-phase system of 0.83 mol/L of the reactant. For the
RuMo/ZrO2 catalyst, the effect of Mo/Ru ratio (01.5) on the activity and selectivity was investigated. The catalyst with the Mo/Ru
ratio of 0.2 exhibited the highest activity for propanoic acid conversion. The activity decreased with further increase in the Mo/Ru
ratio. The products included propanol, methane, ethane, propane
and trace amounts of ethanol and acetic acid. Over Ru/ZrO2 catalyst,
C C bond cleavage to methane and ethane was dominant reaction
compared with the HYD of C O bond, while the RuMo/ZrO2 catalyst
favored the HYD of C O bond in propanoic acid.
The Pt nanoparticles protected by polyethyleneimine were used
as catalyst for the conversion of glucose and fructose at 403543 K

50

E. Furimsky / Catalysis Today 217 (2013) 1356

in a subcritical water and H2 pressure of 5 MPa using a batch reactor


[277]. At 403 K, the former could be readily isomerized to fructose. In the temperature range of 483543 K, glucose produced
1,2-propanediol, 1,2-hexanediol and ethylene glycol, while fructose yielded 1,2-propanediol, 1,2-hexanediol, and glycerol. Other
catalysts tested included the Pt protected by polyvinylpyrrolidone,
Pt/SiO2 and Pt/Al2 O3 . The Pt protected by polyethyleneimine exhibited superior activity.
Hong et al. [278] studied the effect of support on the activity
of Pt (1 wt.%) catalysts during the conversion of phenol. In this
case, zeolites such as HY, H and HZSM-5 as well as -Al2 O3 and
SiO2 , were compared (473 and 523 K; H2 pressure of 4 MPa; WHSV
of 20 h1 ; 10 wt.% H2 O). For all catalysts, the overall phenol conversion reached almost 100%. However, signicant difference in
product distribution, was observed. Thus, cyclohexane accounted
for more than 90% of all products over Pt/zeolite catalysts compared with less than 3 wt.% over Pt/Al2 O3 and Pt/SiO2 catalysts. For
the latter catalysts, cyclohexanol accounted for almost 95% of the
converted phenol. Small amounts of bicyclics and tricyclic products
were observed over the Pt/zeolite catalyst.
During the HDO of phenol in aqueous medium (below 453 K,
Pd/C catalyst, batch reactor) Zhao et al. [279a,b] observed an
increased yield of cyclohexcanol and decreased yield cyclohexanone with time on stream. This suggests that the latter was an
intermediate for the formation of cyclohexenol. At 453 K, small
amount of cyclohexane was formed after acidifying the solution
with H3 PO4 . There was little evidence of the hydrogenolysis of
phenol to benzene. Similarly, a high selectivity to cyclohexanol
in a neutral aqueous solution was reported over Pt-, Ru-, and
Rh-based catalysts [19]. However, with temperature increase to
473 K of the acidied solution, cyclohexanol was quantitatively
dehydrated to cyclohexene followed by HYD to cyclohexane. It is
therefore believed that the presence of hydronium ions is required
for the dehydration of cyclohexanol, while noble metals facilitate
the HYD function. This concept of catalysis was conrmed using
more complex reactants such as guaiacols and syringols. Thus, guaiacols (4-n-propylguaiacol, 4-allylguaiacol and 4-acetonylguaiacol)
were converted to cycloalkanes (80%), methanol (78%) 1218%
intermediate cycloalcohols or cycloketones (1218%). For these
experiments, the reaction mixture consisted of 5 wt.% Pd/C catalyst
(0.040 g), reactant (0.0106 mol), 0.5 wt.% H3 PO4 in 80 mL of water.
Zhao et al. [280] expanded their study with the aim to develop
a solid acids as the source of hydroniom ions, which are stable
in an aqueous medium at high temperatures. For example, the
Naon polymer used as one of the solid acids was hardly ionized as conrmed by little change in pH with time on stream.
To test the concept, a series of liquid and solid acids in the presence of Pd/C and RANEY Ni catalysts were used for the aqueous
HDO of 4-propylphenol. The results are summarized in Table 24
[280]. A combination of aqueous solutions of either H3 PO4 or
CH3 COOH with Pd/C yielded 84 and 74% of propylcyclohexane,
respectively, while 98% yield of propylcyclohexane was obtained
with both Naon suspension in water and the Naon supported on
SiO2 (13 wt.% of Naon). A combination of either Naon water suspensions or the Naon/SiO2 composite with freshly prepared Ni
catalysts (RANEY Ni) resulted in 100% n-propylcyclohexane yield
compared with 51 and 96% yields for commercial RANEY Ni 2400
and RANEY Ni 4200, respectively. Compared with Naon, zeolites
were poor source of hydronium ions as indicated by rather low
yields of propylcyclohexane [271]. The combination of RANEY Ni
with Naon/SiO2 was used to study the HDO of 2-methoxy-4-npropylphenol. Table 25 [280] shows that temperature of 473 K was
not sufcient for achieving a high conversion.
Ohta et al. [281] prepared series of Pt catalysts (2 wt.% Pt) supported on activated carbon (Norit and Wako), mesoporous carbon,
multi-walled carbon nanotube and carbon black, via impregnation

Table 24
Aqueous-phase HDO of 4-n-propylphenol over Pd and/or Ni based catalysts and
acids (473 K, 4 MPa H2 and 0.5 h).
Catalyst

Acid

Conv. (%)

Cycloalkane select.
(%)

Pd/C
Pd/C
RANEY Ni
RANEY Ni
Pd/C
Pd/C
Pd/C
Pd/C
RANEY Ni
RANEY Ni
RANEY Ni 2400
RANEY Ni 4200
Ni/SiO2
Ni/ASA

H3 PO4
CH3 COOH
H3 PO4
CH3 COOH
Zeolite (H-Beta)
Zeolite (H-Y)
Naon solution
Naon/SiO2
Naon/SiO2
Naon solution
Naon/SiO2
Naon/SiO2
Naon/SiO2
Naon/SiO2

100
100
0
0
100
100
100
100
100
100
51
96
9
37

84
74

1.5
5.2
98
98
99
98
36
64
43
50

Ref. [280].

Table 25
HDO of 2-methoxy-4-n-propylphenol over RANEY Ni + Naon/SiO2 .
T (K)

Conv. (%)

Cycloalkanes

Methanol

473
523
573

1
38
80

89
88
86

9.3
8.4
8.7

Ref. [280].

with aqueous solution of H2 PtCl6 . The catalysts were used for the
HDO of 4-propylphenol in water at 280 C under 4 MPa H2 . The
Pt/ACN exhibited high activity with 97% yield of propylcyclohexane similarly as the Pt catalyst supported on mesoporous carbon
and carbon nano tube. The Pt supported on carbon black was less
active. The Pt catalysts supported on ZrO2 , TiO2 and CeO2 were
moderately active but beside propylcyclohexane, propylbenzene
in 310% yield was also present. Contrary to these observations,
Pt/Al2 O3 catalyst was inactive because of the structural transformation of -Al2 O3 into boehmite. The activity of the Rh, Ru and Pd
catalysts supported on activated carbon was much lower than that
of Pt/C (Table 24).
8.2. Super critical conditions
Under super critical conditions, water is completely miscible
with organic compounds. At the same time, gaseous H2 is completely miscible with SCW as well. This ensures a homogeneity of
reaction streams and an efcient transfer of hydrogen to reactant
molecules.
Duan and Savage [282] studied hydrothermal transformations
of pyridine in an aqueous medium with aim to simulate HPR of
an algae biofeed under SCW conditions. Nitrogen removal from
algae biofeed above critical temperature of water may have some
advantages. For example, large portion of ammonia produced from
nitrogen compounds ends up in aqueous phase rather than in oil
phase. The efciency of upgrading can be enhanced in the presence
of catalyst. In this regard, a series of commercially available catalysts, i.e., 5% Pt/C, 5% Pd/C, 5% Ru/C, 5% Rh/C, 5% Pt/C-sulded, 5%
Pt/Al2 O3 , Mo2 C, MoS2 , PtO2 , Al2 O3 , CoMo/Al2 O3 -sulded and activated carbon, were evaluated for potential applications (380, 400,
420 C, 10150 min, catalyst loading of 50200 wt.%, H2 pressure of
06.9 MPa). The 5% Pt/Al2 O3 catalyst exhibited the highest activity
and stability.
The SCW hydrothermal process, operating in the presence of
the Pt/C catalyst under a high H2 pressure, was developed by
Duan and Savage [283,284a] for upgrading biofeed from liquefaction of a microalgae. The upgraded oil was a freely owing

E. Furimsky / Catalysis Today 217 (2013) 1356

51

liquid compared with a tarry consistency biofeed. The characterization of upgraded bio oil identied 72 compounds which
accounted for almost 70% of the product mixture A series of nalkanes starting at about C9 were dominant species in products.
This conrmed a high level of the HYD of alkenes present in the feed.
For example, phytenes present in the biofeed were converted to
phytane (2,6,10,14-tetramethylhexadecane). Signicant amounts
of alkyl substituted benzenes were formed as well. Derivatives of
piperidine, indole and O-methyloxime, which were present in
the crude material, were not detected in the products after the
supercritical upgrading process. This suggests that the catalyst
and reaction conditions used caused an extensive denitrogenation. The overall content of cholesterol, cholestane, and cholestene
decreased in the treated oils. The only sulfur-containing compound,
such as 1-methyl-2-piperidinethione, detected in the biofeed, was
removed during upgrading. The experimental conditions had a pronounced effect on the products distribution [284b]. Without Pt/C
catalyst, the content of fatty acid in the produced liquids was rather
high. Basic conditions favored the conversion of fatty acids. At the
same time, the relative amount of pentadecane in the product oil
was much higher than that in the crude feed. In another study,
the biofeed from the hydrothermal liquefaction of a microalgae
was upgraded over the Pd (5 wt.%)/carbon catalyst in a SCW at
400 C and 3.4 MPa H2 [283]. The longer reaction times and a lower
feed/catalyst ratio increased the amount of gas and coke. At the
same time, the quality of liquid products improved although their
yield decreased.

respectively. The studies conducted by Li et al. [284b,c] may be


introduced to illustrate the biofeed upgrading in the supercritical methanol. In this case, low boiling biofeed was upgraded over
Pt/Al2 (SiO3 )3 , Pt/C and Pt/MgO catalysts in supercritical methanol
in the presence of H2 . Tang et al. [284d,e] studied the HPR of lignin to
produce liquid fuels over Ru/ZrO2 /SBA-15 catalyst in supercritical
ethanol. These studies were conducted in batch systems.
The objective of the study conducted by Dang et al. [284f]
was the conversion of unstable species present in a biofeed (e.g.,
acids, phenols, aldehydes, sugars) into more stable oxygenates
over bifunctional Pt/SO4 2 ZrO2 /SBA-15 catalyst in supercritical
ethanol. The effect of the initial H2 pressures (0.52.0 MPa),
the ethanol/biofeed ratio (5:1, 3:1, 2:1, 1:1) and temperatures
(260300 C), were investigated. When temperature was increased
from 260 to 300 C, the total reaction pressures ranged between
about 7.011.8 MPa. The H2 pressure increase resulted in a signicant decrease in coke deposition on catalyst surface. Level of
upgrading increased with the increasing ethanol/biofeed ratio.
The study of Wang and Rinaldi [284g] may open a new venue for
development of biocrude upgrading process. In this study, propan2-ol was used as the hydrogen donor in the presence of a RANEY
Ni catalyst under rather mild conditions. Thus, at 160 C, an extensive HYD of phenolic rings to cyclic alcohols could be achieved. In
simultaneous reactions, ketones and aldehydes were converted to
corresponding alcohols. However, catalyst deactivation caused by
the presence of water was quite evident. Nevertheless, a search for
more stable catalysts may improve viability of this method.

9. Hydroprocessing of biofeeds in protic solvents

10. Renery aspects of biofuels production

The homogeneity of biofeed for HPR may be improved by dissolution in polar protic solvents (e.g., methanol, ethanol, diethylene
glycol, etc.). This may be critical for biofeeds derived from municipal
solid waste. The miscibility of alcohols with biofeeds and H2 is signicantly enhanced under supercritical conditions. Consequently,
the rates of HPR reactions are increased. Apparently, adverse effects
on catalyst activity caused by an excessive formation of water
may be anticipated. In addition, esterication of acids present in
biofeeds to various esters may not be avoided.
As indicated in Sections 2 and 3, the information on the production and characterization of the biofeeds derived from sewage
sludge is rather limited. Yet, the conversion of these materials to
biofuels may face signicant challenges, as it was indicated in the
study published by Izhar et al. [55c] who used a conventional sulded NiMoP/Al2 O3 catalyst for the HPR obtained from a sewage
sludge by pyrolysis [55b]. The composition of this biofeed was
described in Section 3. Tests were conducted in a xed bed reactor
at 250 or 350 C, total pressure of 2.0 MPa and the LHSV of 10 h1 .
Because of a high viscosity and/or low pumpability, the biofeed had
to be diluted. A complete miscibility was achieved using 20 wt.% of
biofeed with diethyl glycol, methanol and ethanol. In the case of
xylene, only about two thirds of the 20 wt.% of biofeed were miscible giving about 16 wt.% in the mixture. The insoluble portion was
removed from the feed mixture. The nitrogen in the biofeedxylene
mixture (0.5 wt.%) was completely removed at both 250 and
350 C. At the same time, the oxygen content was reduced from 1.4
to 1.0 and 0.5 wt.% at 250 C and 350 C, respectively. Under identical conditions, only about 11, 27 and 0 wt.% removal of nitrogen was
achieved with the solvents such as methanol, ethanol and diethyl
glycol, respectively. This was attributed to the excessive amount of
water formed in dehydration and HDO reactions.
Most recently, interests in the HPR of biofeeds in supercritical
alcohols have been noted. In this case, methanol and ethanol were
the alcohols of choice. Critical temperatures and critical pressures
of these alcohols are 240 and 243 C as well as 7.9 and 6.3 MPa,

Biofuels may be produced in a stand-alone biorenery providing that the economic parameters are attractive. In this regard, a
number of important facts, i.e., source and location of biomass,
type of biomass, transportation costs, method of biofeed production, availability of hydrogen, scale of biorenery and associated
capital cost, overall emissions, etc., must be taken into consideration. The origin of biofeeds may have a signicant impact on the
HPR downstream options. For example, biofeeds from vegetable
oils can be transported to petroleum renery for further processing
and/or coprocessing. The only stand-alone commercial process for
biofuels production developed by Neste Oil Corp. employs the
biofeed of vegetable oil origin [8]. For a similar purpose, biofeeds
of lignocellulosic origin would require a stabilization step prior to
transportation.
A commercialization stage is approaching the UOP/Eni Econing process developed by Honeywells UOP [285]. This process is
suitable for the production of green diesel and jet fuel [286]. For
the latter, two upgrading steps are necessary. In the rst step, the
biofeed is subjected to HPR to convert glycerides to waxy C16 C18
hydrocarbons. To attain desirable cold ow properties, this product
is treated in the second stage over a catalyst active for HCR and HIS.
Engineering study for 100-million gal/an was performed with the
plant start-up in 2012.
Once in renery, biofeeds may be blended with petroleum feeds
for further HPR to produce commercial fuels. For this purpose, an
optimal volume of biofeed in the blend must be determined to minimize deactivation of HPR catalyst and to ensure desirable quality of
liquid products. In advancing toward environmental renery, modication of conventional processing schemes may be necessary to
enable co-HPR of biofeeds with the feeds of petroleum origin. The
conversion of biofeeds to hydrocarbon fuels may require removal
of signicant amount of oxygen. This may lead to the formation of
an oil phase product and a separate aqueous phase product. Therefore, for some biofeeds, the HPR process must include a system
for the separation of these phases. The importance of HDO during

52

E. Furimsky / Catalysis Today 217 (2013) 1356

Fig. 31. Multi catalysts bed reactor for coprocessing of light gas oil with raw tall
diesel (RTD).
Ref. [285].

the HPR of biofeeds represents the main difference compared with


petroleum feeds as discussed in details in the review published by
Elliot [19].
The existing infrastructure of petroleum reneries may be suitable for coprocessing of biofeeds with petroleum derived feeds.
In this regard, HPR and FCC appear to be the most promising alternatives [168,265b]. Apparently, among different biofeeds,
those derived from vegetable oils may be the most suitable for
coHPR with the feeds of petroleum origin because of their high
energy density, low oxygen content and their liquid form. As
pointed out by Huber and Corma [2,168], blending the vegetable
oil with a petroleum feed dilutes the latter. Consequently, operating parameters must be adjusted to ensure a high rate of HPR

reactions and to maintain a steady catalyst performance for a long


period on stream. However, when the VGO was coprocessed with
a vegetable oils biofeed, little adjustment of operating parameters
was necessary [168].
The process designed by Haldor Topsoe is introduced just to
illustrate the complexity which may be involved during the HPR of
biofeeds [285]. The owsheet of this process (Fig. 31) indicates the
importance of catalyst selection. In total, four different types of catalysts may be needed for co-processing of a conventional LGO with
about 30% of the raw tall diesel (RTD) to produce ULSD. The RTD is
produced by transesterication of tall oil, which is a by-product of
kraft pulping. To avoid corrosion in the units upstream from reactor, RTD is injected together with LGO just before entering the HPR
reactor. All catalysts developed for the process are of a proprietary
nature. However, it may be anticipated that these catalysts exhibit
high HDO and HDS activities. The extent of HDO may determine
the overall hydrogen consumption. For different biofeeds, some
modications of the process (Fig. 31) may be necessary [286,287].
Blending biofuels with petroleum fractions (both after HPR) is
another option available to reneries. Table 26 [15,16] shows the
effect of addition of biodiesel from HPR of rapeseed oil (360 C,
7 MPa and sulded NiMo/Al2 O3 ). The addition of the former to
petroleum diesel shifted distillation curve toward to higher boiling points. The shift was mainly in the middle of distillation curve,
while little effect on the key points of distillation curve, was
observed. Viscosity and density of mixed fuels corresponded to
EN 590 standard requirements except for a little deviation in density for the blend containing 30 wt.% of bio liquid. Flash points
of blends approached that of basic petroleum diesel. Mixed fuels
had benet because of the linear increase in the calculated cetane
index, i.e., from 52 for basic diesel up to 66 units for blend containing 30 wt.% of bio liquid. The increase of cetane number should
exhibit similar trends. The addition of bio-products worsened cold
ow properties of the blends compared with the basic petroleum
diesel fraction. The cloud point and CFPP increase was negligible for
blends containing up to 10 wt.% of bio-liquid while at higher volumes the increase was more evident, i.e., 512 C. However, cold
ow properties could be improved using suitable additives.

11. Observations and future perspectives


For the overall economy of biofuel production, i.e., biomass
planting, maintaining growth, harvesting, transportation and the
feed preparation for biocrude production, must be considered.
The location of the plant for biocrude production and upgrading
requires attention as well. For example, upgrading plant may be

Table 26
Properties of blends of petroleum diesel with biodiesel from HPR of rapeseed oil (360 C and 7 MPa).
Parameter

Density at 15 C, (kg m3 )
Kin. viscosity at 40 C (mm2 /s)
Flash point (closed cup) ( C)
Cetane index
Sulfur (ppm)
Nitrogen (ppm)
Water (ppm)
Low-temperature properties ( C)
Cloud point
CFPP
Pour point
Distillation (recovered vol.%)
at 250 C
at 350 C
95 vol.% recovered at ( C)
Refs. [15,16].

Content of the product (wt.%)

EN 590

10

20

30

829
2.35
72
52
10
37
40

826
2.44
72
54
10
36
41

824
2.50
73
57
10
34
43

820
2.63
74
61
10
32
45

816
2.76
76
66
10
29
48

9
12
12

8
12
12

7
10
9

1
7
9

+3
3
3

48
95
351

45
95
350

42
95
352

36
95
350

27
95
352

820845
2.004.50
>55
>46
<10

<200
<8a
+5 to 20

<65
>85
<360

E. Furimsky / Catalysis Today 217 (2013) 1356

located either as the stand-alone plant at the point of biomass


harvesting or in a separate biorenery. There is also a signicant potential for an integration of the bio fuels production with
petroleum rening. Among several options, HPR has emerged as
the most promising route for biofuels production.
The Nestle process was developed in Finland for biofuels production from biofeeds of the vegetable oils origin. At present time,
this may be the only commercial biofuels production. Apparently,
this process may be operated either as a stand alone plant or in an
integration with conventional petroleum renery. The success in
commercialization of this process was based on the development
of an active HPR catalyst which can exhibit steady performance in
the presence of relatively large quantities of H2 O. Such catalyst was
identied during signicant research efforts in Finland.
Limitations on the use of conventional HPR catalysts have been
noted because they operate in a sulded form and in the presence
of a sulfur donating agent in the biofeed. This may lead to contamination of the products by sulfur which may occur even in the
absence of sulfur in the biofeed because of a sulded form of catalyst employed. It is believed that these problems may be minimized
either by a ne tuning of the HPR operation or employing a blend
of biofeed and a conventional petroleum feed. Otherwise, the production of ultra low sulfur fuels would be affected. Therefore, the
selection of suitable HPR catalyst (e.g., not requiring sulding) may
be key factor for successful operation, particularly for stand alone
plant. Products from the HPR of vegetable oil biofeeds may be used
either directly as green diesel or as a blending component with
diesel fractions derived from petroleum.
Some problems have still to be solved for biofeeds of a lignocellulosic origin before commercialization. It is believed that the
conversion of an algae biomass to biofeeds followed by HPR of the
latter to biofuels is in an early stage of development. In any case,
the production of biofuels should be evaluated in relation to other
sources of fuels (e.g., conventional petroleum, oil shale and CDL).
For overall costs comparison, the efciency of conversion and associated emissions involving full cycle must be considered. It is noted
that for biofuels production, these issues have not been receiving
an adequate attention.
The production of biofuels from lignocellulosic sources requires
more attention because of the presence of high coking propensity
components which make catalyst selection for HPR more challenging. Thus, at least two HPR stages, employing a different catalyst
each, may be necessary to produce commercial fuels. For the
rst (stabilization) stage, catalysts comprising a high HYD activity metals in combination with a low acidity supports were found
to be the most suitable. Because of a high content of phenolic and
aromatic structures in the rst stage products, catalyst for the second stage must exhibit high activity for HDO and HYD. Liquids from
the second stage may be either high in aromatic or in naphthenic
structures. The gasoline fraction derived from the former type of
liquids may be a suitable blending component with similar boiling range fractions of a low aromatics content (e.g., FTS liquids,
low aromatics petroleum liquids, etc.). However, high naphthenic
products usually have a low octane number and cetane number.
They may require additional HPR over catalyst exhibiting high HIS
and ring opening activities. Signicant advances in catalyst development for HPR of lignocellulosic biofeeds were discussed in this
review. Potential of noble metals based catalysts supported on carbon and -Al2 O3 has been noted. These catalysts may exhibit a
desirable activity in the presence of large quantities of H2 O without
requiring presulding compared with conventional HPR catalysts.
Among the biofeeds discussed in this review, the selection of
catalysts for the HPR of biofeeds derived from algae sources is
the most challenging. So far, there is little experience in the HPR
of biofeeds in which nitrogen content may approach 10 wt.%. It
is unlikely that conventional catalysts and noble metals based

53

catalysts can maintain stability for a desirable period of time in


the presence of such amount of nitrogen in these biofeeds. In addition, a specially designed system may be necessary to handle large
amount of ammonia produced. In spite of these anticipated concerns, the catalyst testing and development involving biofeeds of
algae origin has been receiving much less attention compared with
the other types of biofeeds. Apparently, catalyst development for
the HPR of the algae type biofeeds must focus on entirely different catalytic phase which exhibit high stability in the presence of
basic reactants. It is believed that a multi stage HPR process may
be necessary for such applications. For example, a reactor vessel
comprising several sections with different catalyst each, with a provision for a gas removal and/or an addition of make-up hydrogen,
between the sections may be necessary. It is believed that a signicant research and development efforts are needed before the
production of biofuels from the algae biofeeds via HPR approaches
a commercial stage.
Signicant advancements in the understanding of HPR reactions have been made as part of the efforts on development of
the catalysts for HPR of biofeeds. Thus, catalysts which exhibit
desirable activity, not requiring the presence of sulfur in biofeed,
were identied. On the search for an active and selective catalyst,
specially designed experimental conditions used for conversion of
model reactants (e.g., guaiacol, anisol, phenols, carboxylic acids,
esters, etc.) allowed identication and quantication of numerous products and intermediates. Consequently, the most advanced
mechanisms of HPR reactions such as HDO, HYD, transalkylation,
isomerization, etc., have been developed. It is believed that it may
be benecial to revisit other HPR reactions (e.g., HDS and HDN) in
accordance with these new ndings.

References
[1] A. Corma, G.W. Huber, S. Iborra, Chemical Reviews 106 (2006) 4044.
[2] G.W. Huber, A. Corma, Angewandte Chemie International Edition 46 (2007)
7184.
[3] S.N. Naik, P.K. Rout, A.K. Dalai, Renewable and Sustainable Energy Reviews 14
(2010) 578.
[4] A. Demirbas, Applied Energy 88 (2011) 17.
[5] IFP Energies Nouvelles, 2010 Activity Report, p. 14.
[6] R. Mukhopadhyay, Chemical and Engineering News 89/33 (2011) 10.
[7] J. Van Gerpen, Fuel Processing Technology 86 (2005) 1097.
[8] S. Mikkonen, Hydrocarbon Processing 87 (2008) 63.
[9] T.F. Riesing, http://www.oakhavenpc.org/cultivating algae.htm
[10] European Commission, Directive 2003/30/EC, J Eur Union L 123/42, Brussels,
2003.
[11] S. Shelley, Chemical Engineering (June) (2009) 16.
[12] Oilseeds: January 2009 World Market and Trade Monthly Circular,
&HASCHAR; FOP 1-09, USDA.
[13] M. Eisenmenger, N.T. Dunford, F. Eller, S. Taylor, J. Martinez, Journal of the
American Oil Chemists Society 83 (2006) 863.
[14] L. Li, E. Coppola, J. Rine, J.L. Miller, D. Walker, Energy & Fuels 24 (2010) 1305.
[15] P. Simacek, D. Kubicka, G. Sebor, M. Pospisil, Fuel 89 (2010) 611.
[16] P. Simacek, D. Kubicka, G. Sebor, M. Pospisil, Fuel 88 (2009) 456.
[17] H. Heinemann, Petroleum Rener 33 (1954) 161.
[18] D.C. Elliott, E.G. Baker, Energy from Biomass and Waste X, 1986, p. 765.
[19] D.C. Elliot, Energy & Fuels 21 (2007) 1792.
[20] A.A. Peterson, F. Vogel, R.P. Lachance, M. Froling, M.J. Antal Jr., J.W. Tester,
Energy & Environmental Science 1 (2008) 32.
[21] F. Demirbas, Energy Conversion and Management 41 (2000) 633.
[22] T.V. Choudhary, C.B. Phillips, Applied Catalysis 397 (2011) 1.
[23] B. Zhang, M. von Keitz, K. Valentas, Applied Biochemistry and Biotechnology
147 (2008) 143.
[24] F. Goudriaan, Proc. 4th European Motor Biofuels Forum, 2426 November,
Berlin, 2003.
[25] F. Goudriaan, D.G.R. Peferoen, Chemical Engineering Science 45 (1990) 2729.
[26] D. Maldas, N. Shiraishi, Biomass and Bioenergy 12 (1997) 273.
[27] S. Cheng, I. Dcruz, M. Wang, M. Leitch, C. Xu, Energy & Fuels 24 (2010) 4659.
[28] S. Bensaid, R. Conti, D. Fino, Fuel 94 (2012) 324.
[29] T. Komanoya, H. Kobayashi, K. Hara, W.-J. Chun, A. Fukuoka, Applied Catalysis
407 (2011) 188.
[30] A.H. Stiller, D.B. Dadyburjor, J.P. Wann, D. Tian, J.W. Zondlo, Fuel Processing
Technology 49 (1996) 167.
[31] Y. Matsumura, H. Nonaka, H. Yokura, A. Tsutsumi, K. Yoshida, Fuel 78 (1999)
1049.

54

E. Furimsky / Catalysis Today 217 (2013) 1356

[32] (a) B.A. Akash, C.B. Muchmore, S.B. Lalvani, Fuel Processing Technology 37
(1993) 203;
(b) H. Ramsurn, R.B. Gupta, Energy & Fuels 26 (2012) 2365.
[33] (a) D. Mohan, C.U. Pittman Jr., P.H. Steele, Energy & Fuels 20 (2006) 848;
(b) A.V. Bridgwater, Biomass & Bioenergy 38 (2012) 68.
[34] A.V. Bridgwater, G.V.C. Peacocke, Renewable and Sustainable Energy Reviews
4 (2000) 1.
[35] R. Maggi, B. Delmon, Fuel 73 (1994) 671.
[36] H.-W. Chen, Q.-H. Song, B. Liao, Q.-X. Guo, Energy & Fuels 25 (2011) 4655.
[37] C.A. Mullen, G.D. Strahan, A.A. Boateng, Energy & Fuels 23 (2009) 2707.
[38] (a) E. Furimsky, Applied Catalysis 199 (2000) 147;
(b) E.F. Iliopoulou, S.D. Stefanidis, K.G. Kalogiannis, A. Delimitis, A.A. Lappas,
K.S. Triantafyllidis, Applied Catalysis B: Environmental 127 (2012) 281;
(c) A. Gopalratnam, Chemical Engineering Progress (March) (2011) 35.
[39] J.K. Pittman, A.P. Dean, O. Osundeko, Bioresource Technology 102 (2010) 17.
[40] (a) D. Lozowski, Chemical Engineering (August) (2011) 14;
(b) E.P. Knoshaug, A. Darzins, Chemical Engineering Progress (March) (2011)
37;
(c) K.-C. Cheng, K.L. Ogden, Chemical Engineering Progress (March) (2011) 42.
[41] S. Amin, Energy Conversion and Management 50 (2009) 1834.
[42] H. Kanda, P. Li, Fuel 90 (2011) 1264.
[43] (a) N.H. Tran, J.R. Bartlett, G.S.K. Kannangara, A.S. Milev, H. Volk, M.A. Wilson,
Fuel 89 (2010) 265;
(b) H. Kanda, P. Li, T. Yoshimura, S. Okada, Fuel 105 (2013) 535.
[44] U. Jena, K.C. Das, Energy & Fuels 25 (2011) 5472.
[45] (a) L. Garcia Alba, C. Torri, C. Samor, J. van der Spek, D. Fabbri, S.R.A. Kersten,
D.W.F. Brilman, Energy & Fuels 26 (2012) 642;
(b) C. Torri, L. Garcia Alba, C. Samor, D. Fabbri, D.W.F. Brilman, Energy & Fuels
26 (2012) 658.
[46] A.B. Ross, P. Biller, M.L. Kubacki, H. Li, A. Lea-Langton, J.M. Jones, Fuel 89 (2010)
2234.
[47] T. Minowa, S. Yokoyama, M. Kishimoto, T. Okakura, Fuel 74 (1995) 1735.
[48] Y. Dote, S. Sawayama, S. Inoue, T. Minowa, S. Yokoyama, Fuel 73 (1994) 1855.
[49] Y. Dote, S. Inoue, T. Ogi, S. Yokoyama, Biomass and Bioenergy 11 (1996) 491.
[50] P.J. Valdez, J.G. Dickinson, P.E. Savage, Energy & Fuels 25 (2011) 3235.
[51] D. Zhou, L. Zhang, S. Zhang, H. Fu, J. Chen, Energy & Fuels 24 (2010) 4054.
[52] T. Matsui, A. Nishihara, C. Ueda, M. Ohtsuki, N. Ikenaga, T. Suzuki, Fuel 76
(1997) 1043.
[53] P. Duan, P.E. Savage, Industrial and Engineering Chemistry Research 50 (2011)
52.
[54] N. Ikenaga, C. Ueda, T. Matsui, M. Ohtsuki, T. Suzuki, Energy & Fuels 15 (2001)
350.
[55] (a) X. Pei, X. Yuan, G. Zeng, H. Huang, J. Wang, H. Li, H. Zhu, Fuel Processing
Technology 93 (2011) 35;
(b) J.P. Cao, X.Y. Zhao, X.B. Xiao, S.Y. Zhang, K. Sato, Y. Ogawa, X.Y. Wei, T.
Takarada, Bioresource Technology 102 (2011) 2009;
(c) S. Izhar, S. Uehara, N. Yoshida, Y. Yamamoto, T. Morioka, M. Nagai, Fuel
Processing Technology 101 (2012) 10;
(d) A.I. Rushdi, K.F. Al-Mutlaq, S.K. Sasmal, B.R.T. Simoneit, Fuel 103 (2013)
970.
[56] E. Furimsky, Energy Sources, Part A 30 (2008) 119.
[57] J. Manganaro, B. Chen, J. Adeosun, S. Lakhapatri, D. Favetta, A. Lawal, R. Farrauto, L. Dorazio, D.J. Rosse, Energy & Fuels 25 (2011) 2711.
[58] M. Martin, I.E. Grossmann, Industrial and Engineering Chemistry Research 50
(2011) 15485.
[59] A. de Klerk, E. Furimsky, Catalysis in the Rening of FischerTropsch Liquids,
RSC Publishers, Cambridge, 2010.
[60] A.T. Bell, Catalysis Review Science and Engineering 23 (1981) 203.
[61] R.B. Anderson, The FischerTropsch Synthesis, Academic Press, Orlando, FL,
1984.
[62] M.E. Dry, in: J.R. Anderson, M. Boudart (Eds.), Catalysis Science and Technology, vol. 1, Springer, Berlin, 1981, p. 159.
[63] E. Furimsky, Applied Catalysis 171 (1998) 177.
[64] Y. Chen, S. Guo, C. Wang, F. Yang, Z. Yang, Fuel 93 (2012) 528.

[65] P. Simcek,
D. Kubicka, I. Kubickov, F. Homola, M. Pospsil, J. Chudoba, Fuel
90 (2011) 2473.
[66] R. Maggi, B. Delmon, Biomass & Bioenergy 9 (1994) 245.
[67] R. Maggi, in: B. Delmon, A.V. Brighwater (Eds.), Advances in Thermochemical
Biomass Conversion, vol. 2, Blackie, London, 1992, p. 1086.
[68] P.T. Patil, U. Armbruster, M. Richter, A. Martin, Energy & Fuels 25 (2011)
4713.
[69] A.L. Jongerius, R. Jastrzebski, P.C.A. Bruijnincx, B.M. Weckhuysen, Journal of
Catalysis 285 (2012) 315.
[70] P. De Wild, R. Van der Laan, A. Kloekhorst, E. Heeres, Environmental Progress
& Sustainable Energy 28 (2009) 461.
[71] D.C. Elliott, T.R. Hart, G.G. Neuenschwander, L.J. Rotness, A.H. Zacher, Environmental Progress & Sustainable Energy 28 (2009) 441.
[72] E.D. Christensen, G.M. Chupka, J. Luecke, T. Smurthwaite, T.L. Alleman,
K. Iisa, J.A. Franz, D.C. Elliott, R.L. McCormick, Energy & Fuels 25 (2011)
5462.
[73] D.R. Vardon, B.K. Sharma, J. Scott, G. Yu, Z. Wang, L. Schideman, Y. Zhang, T.J.
Strathmann, Bioresource Technology 102 (2011) 8295.
[74] K. Anastasakis, A.B. Ross, Bioresource Technology 102 (2011) 4876.
[75] T.M. Brown, P. Duan, P.E. Savage, Energy & Fuels 24 (2010) 3639.
[76] P. Biller, A.B. Ross, Bioresource Technology 102 (2011) 215.
[77] F.E. Massoth, Advances in Catalysis 27 (1978) 265.

[78] P. Ratnasamy, S. Sivashanker, Catalysis Review Science and Engineering 22


(1980) 401.
[79] T. Ho, Catalysis Review Science and Engineering 30 (1988) 117.
[80] E. Furimsky, F.E. Massoth, Catalysis Review Science and Engineering 47 (2005)
297.
[81] R. Prins, Advances in Catalysis 46 (2001) 399.
[82] A. Stanislaus, B.H. Cooper, Catalysis Review Science and Engineering 36 (1994)
75.
[83] H. Topsoe, B.S. Clausen, F.E. Massoth, in: J. Anderson, M. Boudart (Eds.), Catalysis, Science and Technology, vol. 11, Springer, 1996, p. 1.
[84] E. Furimsky, Studies in Surface Science and Catalysis 169 (2007) 1.
[85] E. Furimsky, Applied Catalysis 240 (2003) 1.
[86] D.L. Trimm, Design of Industrial Catalysts, Elsevier, Amsterdam, 1980.
[87] D.L. Trimm, A. Stanislaus, Applied Catalysis 21 (1986) 215.
[88] V.J. Lostaglio, J.D. Carruthers, Chemical Engineering Progress (March) (1986)
46.
[89] P. Dufresne, N. Brahma, F. Labruyere, M. Lacroix, M. Breysse, Catalysis Today
29 (1996) 251.
[90] E. Furimsky, F.E. Massoth, Catalysis Today 52 (1999) 381.
[91] A.C. Jacobsen, B.H. Cooper, P.N. Hannerup, Proc. 12th Petrol. World Congr.,
vol. 4, 1987, p. 97.
[92] M. Breysse, E. Furimsky, S. Kasztelan, G. Perot, Catalysis Review Science and
Engineering 44 (2002) 651.
[93] A. Mara, A. Stanislaus, E. Furimsky, Catalysis Review Science and Engineering
52 (2010) 204.
[94] E. Furimsky, Catalysis Today 30 (1996) 223.
[95] A. Prakash, Rening Processes Handbook, Elsevier, Amsterdam, 2002.
[96] P. Zeuthen, K.G. Knudsen, D.D. Whitehurst, Catalysis Today 65 (2001) 307.
[97] J. Miciukiewicz, J. Zmierczak, F.E. Massoth, Proc. 8th Intern. Congr. Catal., vol.
2, Berlin, 1984, p. II-671.
[98] F.E. Massoth, J. Miciukiewicz, Journal of Catalysis 101 (1986) 505.
[99] S. Helveg, J.V. Lauritsen, E. Lgsgaard, I. Stensgaard, J.K. Nrskov, B.S. Clausen,
H. Topse, F. Besenbacher, Physical Review Letters 84 (2000) 951.
[100] M. Breysse, B.A. Bennett, D. Chadwick, M. Vrinat, Bulletin des Societes Chimiques Belges 90 (1981) 1271.
[101] S.M.A.M. Bouwens, F.B.M. van Zon, M.P. van Dijk, A.M. van der Kraan, V.H.J. de
Beer, J.A.R. van Veen, D.C. Koningsberger, Journal of Physical Chemistry 146
(1994) 375.
[102] B.S. Clausen, H. Topse, R. Candia, J. Villadsen, B. Lengeler, J. Als-Nielsen, F.
Christensen, Journal of Physical Chemistry 85 (1981) 3868.
[103] H. Topse, R. Candia, N.-Y. Topse, B.S. Clausen, Bulletin des Societes Chimiques Belges 93 (1984) 783.
[104] H. Topsoe, Applied Catalysis 322 (2007) 3.
[105] M. Brorson, A. Carlson, H. Topsoe, Catalysis Today 123 (2007) 31.
[106] J.V. Lauritsen, M. Nyberg, R.T. Vang, M.V. Bollinger, B.S. Clausen, H. Topse,
K.W. Jacobsen, E. Lgsgaard, J.K. Nrskov, F. Besenbacher, Nanotechnology
14 (2003) 385.
[107] J.V. Lauritsen, M. Nyberg, J.K. Nrskov, B.S. Clausen, H. Topse, E. Lgsgaard,
F. Besenbacher, Journal of Catalysis 224 (2004) 94.
[108] P.G. Moses, B. Hinnemann, H. Topse, J.K. Nrskov, Journal of Catalysis 248
(2007) 188.
[109] J.V. Lauritsen, M.V. Bollinger, E. Lgsgaard, K.W. Jacobsen, J.K. Nrskov,
B.S. Clausen, H. Topse, F. Besenbacher, Journal of Catalysis 221 (2004)
510.
[110] B. Temel, A.K. Tuxen, J. Kibsgaard, N.-Y. Topse, B. Hinnemann, K.G. Knudsen,
H. Topse, J.V. Lauritsen, F. Besenbacher, Journal of Catalysis 271 (2010) 280.
[111] . Logadttir, P.G. Moses, B. Hinnemann, N.Y. Topse, K.G. Knudsen, H. Topse,
J.K. Nrskov, Catalysis Today 111 (2006) 44.
[112] J.V. Lauritsen, S. Helveg, E. Lgsgaard, I. Stensgaard, B.S. Clausen, H. Topse,
F. Besenbacher, Journal of Catalysis 197 (2001) 1.
[113] J. Kibsgaard, A. Tuxen, K.G. Knudsen, M. Brorson, H. Topse, E. Lgsgaard, J.V.
Lauritsen, F. Besenbacher, Journal of Catalysis 272 (2010) 195.
[114] J.V. Lauritsen, J. Kibsgaard, G.H. Olesen, P.G. Moses, B. Hinnemann, S. Helveg,
J.K. Nrskov, B.S. Clausen, H. Topse, E. Lgsgaard, F. Besenbacher, Journal of
Catalysis 249 (2007) 220.
[115] M.V. Bollinger, J.V. Lauritsen, K.W. Jacobsen, J.K. Nrskov, S. Helveg, F. Besenbacher, Physical Review Letters 87 (2001) 196803.
[116] Y.Q. Yang, C.T. Tye, K.J. Smith, Catalysis Communications 9 (2008) 1364.
[117] A. Hrabar, J. Hein, O.Y. Gutirrez, J.A. Lercher, Journal of Catalysis 281 (2011)
325.
[118] X.D. Wen, Z. Cao, Y.W. Li, J. Wang, H. Jiao, Journal of Physical Chemistry B 110
(2006) 23860.
[119] R.R. Chianelli, G. Berhault, Catalysis Today 53 (1999) 357.
[120] G. Berhault, A. Mehta, A.C. Pavel, J. Yang, L. Rendon, M.J. Yacaman, L.C. Araiza,
A.D. Moller, R.R. Chianelli, Journal of Catalysis 198 (2001) 9.
[121] S. Kasztelan, Comptes Rendus de lAcadmie des Sciences. Srie II 307 (1988)
727.
[122] J. Kibsgaard, J.V. Laurentsen, E. Loegsgaard, B.S. Clausen, H. Topsoe, F. Besenbacher, Journal of the American Chemical Society 128 (2006) 13950.
[123] A. Tuxen, H. Gbel, B. Hinnemann, Z. Li, K.G. Knudsen, H. Topse, J.V. Lauritsen,
F. Besenbacher, Journal of Catalysis 281 (2011) 345.
[124] K. Al Dalama, A. Stanislaus, E. Furimsky, Petroleum Science and Technology
2013, in press.
[125] M. Badawi, J.F. Paul, S. Cristol, E. Payen, Y. Romero, F. Richard, S. Brunet, D.
Lambert, X. Portier, A. Popov, E. Kondratieva, J.M. Goupil, J. El Fallah, J.P. Gilson,
L. Mariey, A. Travert, F. Maug, Journal of Catalysis 282 (2011) 155.

E. Furimsky / Catalysis Today 217 (2013) 1356


[126] O.I. Senol, E.M. Ryymin, T.R. Viljava, A.O.I. Krause, Journal of Molecular Catalysis A 277 (2007) 107.
[127] O.I. Senol, E.M. Ryymin, T.R. Viljava, A.O.I. Krause, Journal of Molecular Catalysis A 268 (2007) 1.
[128] T.R. Viljava, R.S. Komulainen, A.O.I. Krause, Catalysis Today 60 (2000) 83.
[129] E.M. Ryymin, M.L. Honkela, T.R. Viljava, A.O.I. Krause, Applied Catalysis 358
(2009) 42.
[130] O. Senol, T. Viljava, A. Krause, Catalysis Today 106 (2005) 186.
[131] O.I. Senol, T.-R. Viljava, A.O.I. Krause, Applied Catalysis 326 (2007) 236.
[132] M.H. Yang, P. Grange, B. Delmon, Applied Catalysis 154 (1997) L7.
[133] E. Laurent, A. Centeno, B. Delmon, Studies in Surface Science and Catalysis 88
(1994) 573.
[134] B. Rozmyslowicz, P. Mki-Arvela, S. Lestari, O.A. Simakova, K. Ernen, I.
Simakova, D.Y. Murzin, T.O. Salmi, Topics in Catalysis 53 (2010) 1274.
[135] D. Kubicka, J. Horcek, Applied Catalysis 394 (2011) 9.
[136] J. Weitkamp, Industrial and Engineering Chemistry Product Research and
Development 21 (1982) 550.
[137] P. Mki-Arvela, M. Snare, K. Ernen, J.D. Myllyoja, Yu Murzin, Fuel 87 (2008)
3543.
[138] K. Miga, K. Stanczyk, C. Sayag, D. Brodzki, G. Djega-Mariadassou, Journal of
Catalysis 183 (1999) 63.
[139] W. Zhang, Y. Zhang, L. Zhao, W. Wei, Energy & Fuels 24 (2010) 2052.
[140] H. Tominaga, M. Nagai, Applied Catalysis 389 (2010) 195.
[141] H.Y. Zhao, D. Li, P. Bui, S.T. Oyama, Applied Catalysis 391 (2011) 305.
[142] E. Furimsky, Carbons and Carbon Supported Catalysts in Hydroprocessing,
RSC Publishing, Cambridge, UK, 2008.
[143] R. Orinkov, A. Orink, Fuel 90 (2011) 3123.
[144] L.B. Sun, Z.M. Zong, J.H. Kuo, G.F. Liu, X. Sun, X.Y. Wei, G.J. Zhou, C.W. Lee,
Energy & Fuels 19 (2005) 1.
[145] E. Furimsky, Catalysis Review Science and Engineering 25 (1983) 421.
[146] M. Ferrari, A. Centeno, C. Lahousse, R. Maggi, P. Grange, B. Delmon, American
Chemical Society, Division of Petroleum Chemistry Preprint 43 (1998) 94.
[147] E. Laurent, B. Delmon, Applied Catalysis 109 (1994) 97.
[148] D. Costa, C. Arrouvel, M. Breysse, H. Toulhoat, P. Raybaud, Journal of Catalysis
246 (2007) 325.
[149] T. Nimmanwudipong, R. Runnebaum, D. Block, B.C. Gates, Catalysis Letters
141 (2011) 817.
[150] (a) T. Nimmanwudipong, R.C. Runnebaum, D.E. Block, B.C. Gates, Energy &
Fuels 25 (2011) 3417;
(b) S. Nimmanwudipong, C. Aiden, J. Lu, R.C. Runnebaum, K.C. Brodwater, N.D.
Browning, D.E. Block, B.C. Gates, Catalysis Letters 142 (2012) 1190.
[151] Y.C. Lin, C.L. Li, H.P. Wan, H.T. Lee, C.F. Liu, Energy & Fuels 25 (2011) 890.
[152] (a) C.V. Loricera, B. Pawelec, A. Infantes-Molina, M.C. lvarez-Galvn, R.
R. Nava, J.L.G. Fierro, Catalysis Today 172 (2011) 103;
Huirache-Acuna,
(b) T. Nimmanwudipong, R.C. Runnebaum, S.E. Ebeler, D.E. Block, B.C. Gates,
Catalysis Letters 142 (2012) 151.
[153] R.C. Runnebaum, R.J. Lobo-Lapidus, T. Nimmanwudipong, D.E. Block, B.C.
Gates, Energy & Fuels 25 (2011) 4776.
[154] T. Nimmanwudipong, R.C. Runnebaum, K. Tay, D.E. Block, B.C. Gates, Catalysis
Letters 141 (2011) 1072.
[155] J. Horcek, F. Homola, I. Kubickov, D. Kubicka, Catalysis Today 179 (2012)
191.
[156] M. Kleinert, J.R. Gasson, T. Barth, Journal of Analytical and Applied Pyrolysis
85 (2009) 108.
[157] T. Barth, M. Kleinert, Chemical Engineering and Technology 31 (2008) 773.
[158] E. Furimsky, Erdoel und Kohle 36 (1983) 518.
[159] L.D. Rollman, Journal of Catalysis 46 (1977) 243.
[160] E.O. Odebunmi, D.F. Ollis, Journal of Catalysis 89 (1983) 56.
[161] C.L. Li, Z.R. Xu, Z.A. Cao, B.C. Gates, AIChE Journal 31 (1985) 170.
[162] F.E. Massoth, P. Politzer, M.C. Concha, J.S. Murray, J. Jakowski, J. Simons, Journal
of Physical Chemistry B 110 (2006) 14283.
[163] Y. Romero, F. Richard, S. Brunet, Applied Catalysis B: Environmental 98 (2010)
213.
[164] A. Popov, E. Kondratieva, J.-P. Gilson, L. Mariey, A. Travert, F. Maug, Catalysis
Today 172 (2011) 132.
[165] O. Yepez, Fuel 86 (2007) 1162.
[166] O. Yepez, Fuel 84 (2005) 97.
[167] J.C. Afonso, M. Schmal, J.N. Cardozo, Industrial and Engineering Chemistry
Research 31 (1992) 1045.
[168] G.W. Huber, P. OConnor, A. Corma, Applied Catalysis 329 (2007) 120.
[169] M.R. de Brimont, C. Dupont, A. Daudin, C. Geantet, P. Raybaud, Journal of
Catalysis 286 (2012) 153.
[170] E.-M. Ryymin, M.L. Honkela, T.-R. Viljava, A.O.I. Krause, Applied Catalysis 389
(2010) 114.
[171] O.I. Senol, T.-R. Viljava, A.O.I. Krause, Catalysis Today 100 (2005) 331.
[172] B. Donnis, R. Egeberg, P. Blom, K. Knudsen, Topics in Catalysis 52 (2009) 229.
[173] M.J. Girgis, B.C. Gates, Industrial and Engineering Chemistry Research 33
(1994) 1098.
[174] M.V. Landau, Catalysis Today 36 (1997) 393.
[175] C.S. Korre, M.T. Klein, R.J. Quan, Industrial and Engineering Chemistry
Research 34 (1995) 101.
[176] C. Aubert, R. Durand, P. Geneste, C. Moreau, Journal of Catalysis 112 (1988)
12.

[177] D. Kubicka, P. Simcek,


N. Z ilkov, Topics in Catalysis 52 (2009) 161.
[178] Q. Smejkal, L. Smejkalov, D. Kubicka, Chemical Engineering Journal 146
(2009) 155.

55

[179] A.T. Madsen, E.H. Ahmed, C.H. Christensen, R. Fehrmann, A. Riisager, Fuel 90
(2011) 3433.
[180] L. Boda, G. Onyestyk, H. Solt, F. Lnyi, J. Valyon, A. Thernesz, Applied Catalysis
374 (2010) 158.
[181] M. Snre, I. Kubickov, P. Mki-Arvela, D. Chichova, K. Ernen, D.Y. Murzin,
Fuel 87 (2008) 933.
[182] I. Kubickova, M. Snare, P. Mki-Arvela, K. Ernen, D.Y. Murzin, Catalysis Today
106 (2005) 197.
[183] (a) J.G. Immer, M.J. Kelly, H.H. Lamb, Applied Catalysis 375 (2010) 134;
(b) B. Rozmyslowicz, P. Mki-Arvela, A. Tokarev, A.-R. Leino, K. Ernen, D.Y.
Murzin, Industrial and Engineering Chemistry Research 51 (2012) 8922.
[184] J.F. Allain, P. Magnoux, P. Schultz, M. Guisnet, Applied Catalysis 152 (1997)
221.
[185] E. Laurent, B. Delmon, Applied Catalysis 109 (1994) 77.
[186] E. Laurent, B. Delmon, Studies in Surface Science and Catalysis 88 (1994) 460.
[187] A. Centeno, E. Laurent, B. Delmon, Journal of Catalysis 154 (1995) 288.
[188] R.C. Runnebaum, T. Nimmanwudipong, R.R. Limbo, D.E. Block, B.C. Gates,
Catalysis Letters 142 (2012) 745.
[189] S.J. Hurff, M.T. Klein, Industrial and Engineering Chemistry Fundamentals 22
(1983) 426.
[190] F.P. Petrocelli, M.T. Klein, Fuel Science & Technology International 5 (1987)
25.
[191] J. Shabtai, N.K. Nag, F.E. Massoth, Journal of Catalysis 104 (1987) 413.
[192] L. Artok, O. Erbatur, H.H. Schobert, Fuel Processing Technology 47 (1996) 153.
[193] S.R. Kirby, C. Song, H.H. Schobert, Catalysis Today 31 (1996) 121.
[194] P.T. Do, M. Chiappero, L.L. Lobban, D.E. Resasco, Catalysis Letters 130 (2009)
9.
[195] (a) M. Snare, I. Kubickova, P. Maki-Arvela, K. Eranen, J. Warna, D.Y. Murzin,
Chemical Engineering Journal 134 (2007) 29;
(b) J.G. Dickinson, J.T. Poberezny, P.E. Savage, Applied Catalysis B: Environmental 123/124 (2012) 357.
[196] E.G. Baker, D.C. Elliott, in: A.V. Bridgwater, J.L. Kuster (Eds.), Research in
Biomass Conversion, Elsevier, London, 1988, p. 883.
[197] R.J. French, J. Stunkel, R.M. Baldwin, Energy & Fuels 25 (2011) 3266.
[198] J.D. Rocha, C.A. Luengo, C.E. Snape, Renewable Energy 9 (1996) 950.
[199] Y.-H.E. Sheu, C.V. Philip, R.G. Anthony, E.J. Soltes, Journal of Chromatographic
Science 22 (1984) 497.
[200] Y.-H.E. Sheu, R.G. Anthony, E.J. Soltes, Fuel Processing Technology 19 (1988)
31.
[201] Z. Su-Ping, Y. Yong-Jie, R. Zhengwei, L. Tingchen, Energy Source 25 (2003) 57.
[202] P.M. Mortensen, J.-D. Grunwaldt, P.A. Jensen, K.G. Knudsen, A.D. Jensen,
Applied Catalysis 407 (2011) 1.
[203] I. Sebos, A. Matsoukas, V. Apostolopoulos, N. Papayannakos, Fuel 88 (2009)
145.
[204] P. Maki-Arvela, B. Rozmyslowicz, S. Lestari, O. Simakova, K. Ernen, T. Salmi,
D.Y. Murzin, Energy & Fuels 25 (2011) 2815.
[205] (a) W. Charusiri, T. Vitidsant, Energy & Fuels 19 (2005) 1783;
(b) R.K. Sharma, M. Anand, B.S. Rana, R. Kumar, S.A. Farooqui, M.G. Sibi, A.K.
Sinha, Catalysis Today 198 (2012) 314;
(c) M. Anand, A.K. Sinha, Bioresource Technology 126 (2012) 148;
(d) S. Kovcs, T. Kasza, A. Thernesz, I. Wlhn Horvth, J. Hancsk, Chemical
Engineering Journal 176/177 (2011) 237.
[206] (a) J.C. Hicks, The Journal of Physical Chemistry Letters 18 (2011) 2280;
(b) D.C. Elliott, T.R. Hart, G.G. Neuenschwander, L.J. Rotness, M.V. Olarte, A.H.
Zacher, Y. Solantausta, Energy & Fuels 26 (2012) 3891.
[207] D.C. Elliott, G.G. Neuenschwander, in: A.V. Bridgwater, D.G.B. Boocock (Eds.),
Developments in Thermochemical Biomass Conversion, vol. 1, Blackie Academic and Professional, London, 1996, p. 611.
[208] P.M. Train, M.T. Klein, Fuel Science & Technology 9 (1991) 193.
[209] M.A. Ratcliff, D.K. Johnson, F.L. Posey, H.L. Chum, Applied Biochemistry and
Biotechnology 17 (1988) 151.
[210] A. Oasmaa, R. Alen, D. Meier, Bioresource Technology 45 (1993) 189.
[211] J.E. Miller, L. Evans, A. Littlewolf, D.E. Trudell, Fuel 78 (1999) 1363.
[212] D. Genuit, P. Afanasiev, M. Vrinat, Journal of Catalysis 235 (2005) 302.
[213] (a) V.N. Bui, D. Laurenti, P. Afanasiev, C. Geantet, Applied Catalysis B: Environmental 101 (2011) 239;
(b) B. Yoosuk, D. Tumnantong, P. Pattarapan, Chemical Engineering Science
79 (2012) 1.
[214] W. Wang, Y. Yang, H. Luo, H. Peng, F. Wang, Industrial and Engineering Chemistry Research 50 (2011) 10936.
[215] W. Wang, Y. Yang, H. Luo, W. Liu, Catalysis Communications 11 (2010) 803.
[216] W. Wang, Y. Yang, H. Luo, T. Hu, W. Liu, Catalysis Communications 12 (2011)
436.
[217] W.Y. Wang, Y.Q. Yang, J.G. Bao, H.A. Luo, Catalysis Communications 11 (2009)
100.
[218] W.Y. Wang, Y.Q. Yang, H.A. Luo, W.Y. Liu, Reaction Kinetics, Mechanisms, and
Catalysis 101 (2010) 105.
[219] (a) V.N. Bui, D. Laurenti, P. Delichre, C. Geantet, Applied Catalysis B: Environmental 101 (2011) 246;
(b) P.E. Ruiz, B.G. Frederick, W.J. De Sisto, R.N. Austin, L.R. Radovic, K. Leiva, R.
Garca, N. Escalona, M.C. Wheeler, Catalysis Communications 27 (2012) 44.
[220] X. Zhu, L.L. Lobban, R.G. Mallinson, D.E. Resasco, Journal of Catalysis 281 (2011)
21.
[221] (a) V.A. Yakovlev, S.A. Khromova, O.V. Sherstyuk, V.O. Dundich, D.Y. Ermakov,
V.M. Novopashina, M.Y. Lebedev, O. Bulavchenko, V.N. Parmon, Catalysis
Today 144 (2009) 362;

56

[222]
[223]
[224]

[225]
[226]
[227]
[228]
[229]
[230]
[231]
[232]

[233]

[234]

[235]
[236]
[237]
[238]
[239]
[240]
[241]
[242]
[243]
[244]
[245]
[246]
[247]
[248]

[249]

[250]
[251]
[252]
[253]
[254]

[255]
[256]
[257]
[258]

E. Furimsky / Catalysis Today 217 (2013) 1356


(b) A.R. Ardiyanti, S.A. Khromova, R.H. Venderbosch, V.A. Yakovlev, H.J. Heeres,
Applied Catalysis B: Environmental 117/118 (2012) 105.
C.R. Lee, J.S. Yoon, Y.-W. Suh, J.-W. Choi, J.-M. Ha, D.J. Suh, Y.-K. Park, Catalysis
Communications 17 (2011) 54.
M.A. Gonzalez-Borja, D.E. Resasco, Energy & Fuels 25 (2011) 4155.
(a) A. Gutierrez, R.K. Kaila, M.L. Honkela, R. Slioor, A.O.I. Krause, Catalysis
Today 147 (2009) 239;
(b) C. Liu, Z. Shao, Z. Xiao, C.T. Williams, C. Liang, Energy & Fuels 26 (2012)
4205.
J. Wildschut, F.M. Mahfud, R.H. Venderbosch, H.J. Heeres, Industrial and Engineering Chemistry Research 48 (2009) 10324.
J. Wildschut, F.M. Mahfud, I.M. Cabrera, R.H. Venderbosch, H.J. Heeres, Energy
& Environmental Science 3 (2010) 962.
J. Wildschut, I. Melin-Cabrera, H.J. Heeres, Applied Catalysis B: Environmental 99 (2010) 298.
A.R. Ardiyantia, A. Gutierrez, M.L. Honkela, A.O.I. Krause, H.J. Heeres, Applied
Catalysis 407 (2011) 56.
P.E. Ruiz, K. Leiva, R. Garcia, P. Reyes, J.L.G. Fierro, N. Escalona, Applied Catalysis 384 (2010) 78.
K. Li, R. Wang, J. Chen, Energy & Fuels 25 (2011) 854.
C. Seplveda, K. Leiva, R. Garca, L.R. Radovic, C.I.T. Ghampson, W.J. DeSisto,
J.L. Garca Fierro, N. Escalona, Catalysis Today 172 (2011) 232.
(a) I.T. Ghampson, C. Seplveda, R. Garcia, B.G. Frederick, M.C. Wheeler, N.
Escalona, W.J. DeSisto, Applied Catalysis 413/414 (2011) 78;
(b) I.T. Ghampson, C. Seplveda, R. Garcia, J.L. Garca Fierro, N. Escalona, W.J.
DeSisto, Applied Catalysis 435/436 (2012) 51;
(c) I.T. Ghampson, C. Seplveda, R. Garcia, L.R. Radovic, J.L. Garca Fierro, W.J.
DeSisto, N. Escalona, Applied Catalysis 439/440 (2012) 111.
(a) M.V. Bykova, D.Y. Ermakov, V.V. Kaichev, O.A. Bulavchenko, A.A. Saraev,
M.Y. Lebedev, V.A. Yakovlev, Applied Catalysis B: Environmental 113/114
(2012) 296;
(b) M. Al-Sabawi, J. Chen, Energy & Fuels 26 (2012) 5355;

D. Kallo, Catalysis Com(c) G. Onyestyk, S. Harnos, A. Kaszonyi, M. Stolcova,


munications 27 (2012) 159.
(a) R.N. Olcese, M. Bettahar, D. Petitjean, B. Malaman, F. Giovanella, A. Dufour,
Applied Catalysis B: Environmental 115/116 (2012) 115;
(b) A. Ausavasukhi, Y. Huang, A.T. To, T. Sooknoi, D.E. Resasco, Journal of
Catalysis 290 (2012) 90.
G.N. Da Rocha Filho, D. Brodzki, G. Djega-Mariadassou, Fuel 72 (1993) 543.
J. Gusmao, D. Brodzki, G. Djega-Mariadassou, R. Frety, Catalysis Today 5 (1989)
533.
S. Bezergianni, A. Dimitriadis, A. Kalogianni, P.A. Pilavachi, Bioresource Technology 101 (2010) 6651.
S. Bezergianni, A. Dimitriadis, T. Sfetsas, A. Kalogianni, Bioresource Technology 101 (2010) 7658.
S. Bezergianni, S. Voutetakis, A. Kalogianni, Industrial and Engineering Chemistry Research 48 (2009) 8402.
S. Bezergianni, A. Kalogianni, A. Dimitriadis, Fuel 93 (2012) 638.
M. Krar, S. Kovacs, J. Hancsok, Bioresource Technology 101 (2010) 9287.
P. Priecel, L. Capek, D. Kubicka, F. Homola, P. Rysnek, M. Pouzar, Catalysis
Today 176 (2011) 409.
D. Kubicka, L. Kaluza, Applied Catalysis 372 (2010) 199.
P. Priecel, D. Kubicka, L. Capek, Z. Bastl, P. Rysnek, Applied Catalysis 397
(2011) 127.
D. Kubicka, M. Bejblova, J. Vlk, Topics in Catalysis 53 (2010) 168.
M. Toba, Y. Abe, H. Kuramochi, M. Osako, T. Mochizuki, Y. Yoshimura, Catalysis
Today 164 (2011) 533.
Y. Liu, R. Sotelo-Boys, K. Murata, T. Minowa, K. Sakanishi, Energy & Fuels 25
(2011) 4675.
(a) K. Murata, Y. Liu, M. Inaba, I. Takahara, Energy & Fuels 24 (2010) 2404;
(b) D. Verma, R. Kumar, B.S. Rana, A.K. Sinha, Energy & Environmental Science
4 (2011) 1667.
(a) R. Nava, B. Pawelec, P. Castano, M.C. Alvarez-Galvan, C.V. Loricera, J.L.G.
Fierro, Applied Catalysis B: Environmental 92 (2009) 154;
(b) S. Gong, A. Shinozaki, M. Shi, E.W. Qian, Energy & Fuels 26 (2012) 2394.
B. Veriansyah, J.Y. Han, S.K. Kim, S.-A. Hong, Y.J. Kim, J.S. Lim, Y.-W. Shu, S.-G.
Oh, J. Kim, Fuel 94 (2012) 578.
M. Arend, T. Nonnen, W.F. Hoelderich, J. Fischer, J. Groos, Applied Catalysis
399 (2011) 198.
I. Simakova, O. Simakova, P. Mki-Arvela, D.Y. Murzin, Catalysis Today 150
(2010) 28.
S. Lestari, I. Simakova, A. Tokarev, P. Mki-Arvela, K. Ernen, D.Y. Murzin,
Catalysis Letters 122 (2008) 247.
S. Lestari, P. Mki-Arvela, H. Bernas, O. Simakova, R. Sjholm, J. Beltramini,
G.Q. Max Lu, J. Myllyoja, I. Simakova, D.Y. Murzin, Energy & Fuels 23 (2009)
3842.
S. Lestari, P. Mki-Arvela, I. Simakova, J. Beltramini, G.Q.M. Lu, D.Y. Murzin,
Catalysis Letters 130 (2009) 48.
S. Lestari, P. Mki-Arvela, K. Ernen, J. Beltramini, G.Q.M. Lu, D.Y. Murzin,
Catalysis Letters 134 (2009) 250.
M. Snare, I. Kubickova, P. Maki-Arvela, K. Eranen, D.Y. Murzin, Industrial and
Engineering Chemistry Research 45 (2006) 5708.
(a) P. Maki-Arvela, I. Kubickova, M. Snare, K. Eranen, D.Y. Murzin, Energy &
Fuels 21 (2007) 30;
(b) E. Santillan-Jimenez, T. Morgan, J. Lacny, S. Mohapatra, M. Crocker, Fuel
103 (2013) 1010.

[259] (a) R. Sotelo-Boyas, Y. Liu, T. Minowa, Industrial and Engineering Chemistry


Research 50 (2011) 2791;
(b) S.J. Reaume, N. Ellis, Energy & Fuels 26 (2012) 4514.
[260] O.V. Kikhtyanin, A.E. Rubanov, A.B. Ayupov, G.V. Echevsky, Fuel 89 (2010)
3085.
[261] (a) J. Duan, J. Han, H. Sun, P. Chen, H. Lou, X. Zheng, Catalysis Communications
17 (2011) 76;
(b) H. Wang, S. Yan, S.O. Salley, K.Y.S. Ng, Industrial and Engineering Chemistry
Research 52 (2012) 10066;
(c) G. Onyestyk, S. Harnos, . Szegedi, D. Kall, Fuel 102 (2012) 282;
(d) H. Zuo, Q. Liu, T. Wang, L. Ma, Q. Zhang, Q. Zhang, Energy & Fuels 26 (2012)
3737.
[262] L.W. Hillen, G. Pollard, L.V. Wake, N. White, Biotechnology and Bioengineering
24 (1982) 193.
[263] L.W. Hillen, L.V. Wake, D.R. Warren, Fuel 59 (1980) 446.
[264] L.V. Wake, L.W. Hillen, Journal of Biotechnology and Bioengineering 22 (1980)
1637.
[265] (a) P. Bielansky, A. Weinert, C. Schnberger, A. Reichhold, Fuel Processing
Technology 92 (2011) 2305;
(b) J.A. Melero, M.M. Clavero, G. Caleja, A. Garcia, R. Miravalles, T. Galindo,
Energy & Fuels 24 (2010) 707.
[266] (a) A. Pinheiro, D. Hudebine, N. Dupassieux, C. Geantet, Energy & Fuels 23
(2009) 1007;
(b) V.N. Bui, G. Toussaint, D. Laurenti, C. Mirodatos, C. Geantet, Catalysis Today
142 (2009) 172.
[267] (a) J. Walendziewski, M. Stolarski, R. Luzny, B. Klimek, Fuel Processing Technology 90 (2009) 686;
(b) N. Aribert, A. Daudin, T. Chapus, Preprint Papers American Chemical
Society, Division of Fuel Chemistry 57 (2012) 775.
[268] A.A. Lappas, S. Bezergianni, I.A. Vasalos, Catalysis Today 145 (2009)
55.
[269] R.J. French, J. Hrdlicka, R. Baldwin, Environmental Progress & Sustainable
Energy 29 (2010) 142.
[270] (a) F. de Miguel Mercader, S.R.A. Kersten, N.W.J. Way, C.J. Schaverien, J.A.
Hogendoorn, Applied Catalysis B: Environmental 96 (2010) 57;
(b) R. Tiwari, B.S. Rana, R. Kumar, D. Verma, R. Kumar, R.K. Joshi, M.O. Garg,
A.K. Sinha, Catalysis Communications 12 (2011) 559;
(c) R. Kumar, B.S. Rana, R. Tiwari, D. Verma, R. Kumar, R.K. Joshi, M.O. Garg,
A.K. Sinha, Green Chemistry 12 (2012) 2232.
[271] S. Bezergianni, A. Kalogianni, I.A. Vasalos, Bioresource Technology 100 (2009)
3036.

D. Kubicka, Fuel 89 (2010) 1508;


[272] (a) P. Simcek,
(b) A. Vonortas, Ch. Templis, N. Papayannakos, Energy & Fuels 26 (2012) 3856;
(c) C. Templis, A. Vonortas, I. Sebos, N. Papayannakos, Applied Catalysis B:
Environmental 104 (2011) 324.
[273] (a) L. Yunqui, American Chemical Society, Division of Petroleum Chemistry
Preprint 54 (2009) 100;
(b) S. Bezergianni, A. Dimitriadis, Fuel 103 (2013) 579.
[274] G.W. Huber, J.A. Dumesic, Catalysis Today 111 (2006) 119.
[275] B. Peng, C. Zhao, I. Meja-Centeno, G.A. Fuentes, A. Jentys, J.A. Lercher, Catalysis
Today 183 (2012) 3.
[276] L. Chen, Y. Zhu, H. Zheng, C. Zhang, Y. Li, Applied Catalysis 411/412 (2011)
95.
[277] Y. Kanie, K. Akiyama, M. Iwamoto, Catalysis Today 178 (2011) 58.
[278] D.Y. Hong, S.J. Miller, P.K. Agrawal, C.W. Jones, Chemical Communications 46
(2010) 1038.
[279] (a) C. Zhao, Y. Kou, A.A. Lemonidou, X.B. Li, J.A. Lercher, Angewandte Chemie
International Edition 48 (2009) 3987;
(b) C. Zhao, J. He, A.A. Lemonidou, X. Li, J.A. Lercher, Journal of Catalysis 280
(2011) 8.
[280] C. Zhao, Y. Kou, A.A. Lemonidou, X.B. Li, J.A. Lercher, Chemical Communications 46 (2010) 412.
[281] (a) H. Ohta, H. Kobayashi, K. Hara, A. Fukuoka, Chemical Communications 47
(2011) 12209;
(b) M.V. Olarte, V.M. Lebarbier, H.M. Brown, M. Swita, T. Lemmon, D.C. Elliott,
Preprints of Papers American Chemical Society, Division of Fuel Chemistry 57
(2012) 754.
[282] P. Duan, P.E. Savage, Applied Catalysis B: Environmental 108/109 (2011)
54.
[283] P. Duan, P.E. Savage, Applied Catalysis B: Environmental 104 (2011)
136.
[284] (a) P. Duan, P.E. Savage, Bioresource Technology 102 (2011) 1899;
(b) W. Li, C. Pan, Q. Zhang, Z. Liu, J. Peng, P. Chen, H. Lou, X. Zheng, Bioresource
Technology 102 (2011) 4884;
(c) W. Li, C. Pan, L. Sheng, Z. Liu, P. Chen, H. Lou, X. Zheng, Bioresource Technology 102 (2011) 9223;
(d) Z. Tang, Y. Zhang, Q. Lu, X.F. Zhu, Q.X. Guo, Industrial and Engineering
Chemistry Research 48 (2009) 6923;
(e) Z. Tang, Y. Zhang, Q. Lu, Q.X. Guo, Industrial and Engineering Chemistry
Research 49 (2010) 2040;
(f) Q. Dang, Z. Luo, J. Zhang, J. Wang, W. Chen, Y. Yang, Fuel 103 (2013) 683;
(g) X.Y. Wang, R. Rinaldi, Energy & Environmental Science 5 (2012)
8244.
[285] G. Ondrey, Chemical Engineering 5 (2010) 11.
[286] G. Ondrey, Chemical Engineering 5 (2010) 18.
[287] G. Ondrey, Chemical Engineering 5 (2010) 14.

Das könnte Ihnen auch gefallen