Sunteți pe pagina 1din 16

View Article Online / Journal Homepage / Table of Contents for this issue

Chem Soc Rev

Dynamic Article Links

Cite this: Chem. Soc. Rev., 2011, 40, 52665281

TUTORIAL REVIEW

www.rsc.org/csr

Published on 29 June 2011. Downloaded by Universidad de Malaga on 28/10/2014 15:46:44.

Transformations of biomass-derived platform molecules: from high


added-value chemicals to fuels via aqueous-phase processingw
Juan Carlos Serrano-Ruiz,*a Rafael Luque*b and Antonio Sepulveda-Escribanoa
Received 17th May 2011
DOI: 10.1039/c1cs15131b
Global warming issues and the medium-term depletion of fossil fuel reserves are stimulating researchers
around the world to nd alternative sources of energy and organic carbon. Biomass is considered by experts
the only sustainable source of energy and organic carbon for our industrial society, and it has the potential
to displace petroleum in the production of chemicals and liquid transportation fuels. However, the transition
from a petroleum-based economy to one based on biomass requires new strategies since the petrochemical
technologies, well-developed over the last century, are not valid to process the biomass-derived compounds.
Unlike petroleum feedstocks, biomass derived platform molecules possess a high oxygen content that gives
them low volatility, high solubility in water, high reactivity and low thermal stability, properties that favour
the processing of these resources by catalytic aqueous-phase technologies at moderate temperatures. This
tutorial review is aimed at providing a general overview of processes, technologies and challenges that lie
ahead for a range of dierent aqueous-phase transformations of some of the key biomass-derived platform
molecules into liquid fuels for the transportation sector and related high added value chemicals.
a

Laboratorio de Materiales Avanzados, Universidad de Alicante,


Crta. San Vicente s/n, 03690 San Vicente, Alicante, Spain
b
Departamento de Qumica Organica, Universidad de Cordoba,
Campus Universitario de Rabanales, Edicio Marie Curie (C3),
E-14014 Cordoba, Spain
w In memory of Victor S. Y. Lin, inspiration and friend, who passed
away in May 2010.

1. Introduction

Dr Juan Carlos Serrano-Ruiz


studied Chemistry at the
University of Granada (Spain).
In 2001 he moved to the
University of Alicante (Spain)
where he received a PhD in
Chemistry
and
Materials
Science in 2006 under the
supervision of Profs. Francisco
Rodrguez-Reinoso and Antonio
Sepulveda-Escribano. In January
2008 he was awarded a
Fulbright fellowship to conduct
studies on biomass conversion
to fuels and chemicals by cataJuan Carlos Serrano-Ruiz
lytic approaches in Prof. James
Dumesics research group at the University of Wisconsin
Madison, Madison (USA). In January 2010 he joined the
Advanced Materials Laboratory (LMA) at the University of
Alicante where he has initiated a new research line on catalytic
routes for the conversion of biomass into liquid hydrocarbon fuels.
He is a member of the Spanish Biomass Technology Platform
(BioPlat) and he belongs to the Spanish Biofuels for Transport
Working Group. He is (co-)author of over 30 manuscripts and
book chapters on biomass conversion and catalysis.

Dr Rafael Luque became


Ramon y Cajal Fellow at the
Universidad de Cordoba in
Spain after a three-year postdoctoral stay at the University
of York (UK) in the Green
Chemistry Centre of Excellence directed by Prof. James
Clark. His interests range
from materials science, nanotechnology and heterogeneous
catalysis to biomass valorisation and biofuels. He is a
(co)author of more than 100
peer-reviewed manuscripts, 15
Rafael Luque
book chapters and 2 patents as
well as many invited talks and keynote lectures in International
Conferences worldwide. He recently served as editor of the book
Handbook of biofuels production: processes and technologies
and is currently editing the books Sustainable preparation of
metal nanoparticles from the RSC Green Chemistry book
series, Green Chemistry from Novapublishers and Advances
in biodiesel production from Woodhead Publishing. He
currently sits on the Editorial Board of Current Organic
Synthesis.

5266

Chem. Soc. Rev., 2011, 40, 52665281

The growing scarcity of fossil hydrocarbons, the increase in oil


prices, concerns regarding Green House Gases (GHG) emissions,
and the projected increasing demand for energy in the future are
driving society towards the search for new renewable resources

This journal is

The Royal Society of Chemistry 2011

Published on 29 June 2011. Downloaded by Universidad de Malaga on 28/10/2014 15:46:44.

View Article Online

that can substitute for fossil fuels in the current energy system.
Renewable sources including solar, wind, hydroelectric and
geothermal activity can substitute natural gas and coal in the
production of heat and electricity, while biomass, the only
sustainable source of organic carbon in earth, has been pointed
out as the perfect equivalent to petroleum for the production of
fuels, chemicals and carbon-based materials.1
Unlike petroleum, biomass is renewable, widely available
and, most importantly, some biomass feedstocks including
forestry residues, manure, waste oils and fats can be considered
strictly speaking as residues. Additionally, biomass feedstocks
feature a closed cycle (as compared to the broken cycle in crude
oil) in which the release of CO2 generated in the transformation
to chemical products and/or fuels as well as burning of such
produced fuels is recaptured by the existing plants via photosynthesis during biomass regrowth.
However, the utilization of biomass for the large scale
production of fuels and chemicals is associated with a number
of important issues. Consumption of edible biomass feedstocks (e.g. sugars, starches and vegetable oils) leads to
competition with food for land use, and a strong food versus
fuel debate emerged as a response to the sharp increase in food
prices during 2007 and 2008. The ambitious targets imposed
by governments to displace a signicant fraction of petroleumderived fuels and chemicals within the next years contrast
with the limited availability of agriculture land in many
countries. Additionally, the replacement of classical forest
lands for fuel-crops is creating important deforestation,
global warming and biodiversity threatening concerns in some
developing countries around the world. These issues have
encouraged researchers to develop technologies to process

Antonio Sepulveda-Escribano
is full Professor at the
Inorganic Chemistry Department and member of the
University Materials Institute
of the University of Alicante
(Spain). He received his
PhD from the University of
Alicante in 1989, with a work
on carbon-supported catalysts
for FischerTropsch synthesis.
Then, he worked as postdoctoral
researcher at the Instituto de
Catalisis y Petroleoqumica
Antonio Sepulveda-Escribano (CSIC, Madrid, Spain) and at
the Institut de Recherches sur la
Catalyse (CNRS, Lyon, France) from 1989 to 1993. This year he
went back to the University of Alicante, where he became Assistant
Professor in 1996 and full Professor in 2003. Since 2009 he leads the
Advanced Materials Laboratory research group. He has authored
over 100 publications in peer-reviewed scientic journals, and several
book chapters. His research activities are centred in the elds of
adsorption and catalysis: synthesis, characterization and application
of adsorbent materials (activated carbons, mesoporous silicas), and
supported metal catalysts for dierent applications. Currently, these
applications are mainly related to environmental protection (VOCs
removal, CO2 sequestration) and biomass transformation processes
(hydrogen production and purication).
This journal is

The Royal Society of Chemistry 2011

more abundant and cheaper nonedible biomass (lignocellulosic biomass), thereby permitting sustainable production
of fuels and chemicals without aecting food supplies or
forcing extensive changes in land use.
Lignocellulose is a polymeric material composed of three
primary units: cellulose (a crystalline polymer of glucose
units), hemicellulose (an amorphous polymer of ve dierent
C5 and C6 sugars) and lignin (a three-dimensional polymer of
propyl-phenol that surrounds cellulose and hemicellulose).
This structure provides lignocellulose with a high degree of
chemical complexity and recalcitrance (lignin oers an extra
protection to cellulose and hemicellulose) which increases
processing cost for this resource compared to simpler and
readily degradable edible biomass.
The success of the oil industry during the last century can be
explained in terms of ecient and complete utilization of
petroleum resources by optimised technologies in highlyintegrated facilities. Thus, economic feasibility of petroleum
reneries is achieved through the simultaneous production of
large volumes of low-value transportation fuels along with
lower volumes of more valuable chemicals and carbon-based
products. The enormous and complex petrochemical industry,
however, can be essentially constructed from a few molecules
(benzene, toluene, xylene, ethylene, propylene, 1,3-butadiene
and methanol) which serve as building blocks for the production of a multitude of chemicals and specialty products.
Following a similar approach, a future biomass-based
industry could achieve the desired conversion of a variety of
bio-feedstocks into fuels, valuable chemicals and energy
through the processing of carefully-selected biomass derivatives (the so-called platform molecules) in an integrated facility
denoted as biorenery. The selection of these relevant platform molecules was originally conducted by the USDOE,
from which 12 compounds were highlighted as key starting
materials to focus most research future endeavours.2 Organic
acids such as succinic, itaconic, fumaric, lactic and levulinic,
and polyols including glycerol, sorbitol, and xylitol were some
examples present on this original list. More recently, Bozell
and Petersen revisited this list of compounds and proposed an
updated group of platform molecules (Table 1) which included
many of those originally selected as well as others (e.g.
ethanol, bio-hydrocarbons, furans, etc.) included on the basis
of their current and future potential dictated by a series of
indicators.3
When compared to petroleum feeds, these renewable platform compounds are found to be chemically opposed, with
biomass derivatives presenting a high degree of oxygenated
groups and petroleum feeds being generally unfunctionalized.
This dierent chemical nature forces to design dierent
processing strategies for both resources. For example, since
petroleum feeds are hydrophobic, have low reactivity and high
volatility, high-temperature processes (to activate the relatively inert compounds) and gas phase reactions will be typical
in reneries. In contrast, the high oxygen content of biomass
derivatives makes them water-soluble (they are normally
obtained in the form of dilute aqueous solutions after chemical
and biological processing of biomass sugars), highly-reactive
(with a natural tendency to decompose with temperature)
and low volatile. Such molecules are therefore more suitably
Chem. Soc. Rev., 2011, 40, 52665281

5267

View Article Online

Published on 29 June 2011. Downloaded by Universidad de Malaga on 28/10/2014 15:46:44.

Table 1

Platform molecules: original vs. revisited

Original platform molecules (Werpy and Petersen, 2004)

Revisited platform molecules (Bozell and Petersen, 2010)

Succinic, fumaric and malic acids


3-Hydroxypropanoic acid
Glucaric acid
Itaconic acid
3-Hydroxybutyrolactone
Sugars (sorbitol, xylitol/arabinitol)

Ethanol
Furans (furfural, HMF and 2,5-furandicarboxylic acid)
Glycerol
Bio-hydrocarbons
Lactic acid
Succinic acid
3-Hydroxypropanoic acid
Levulinic acid
Sugars (sorbitol, xylitol)

2,5-Furandicarboxylic acid
Aspartic acid
Glutamic acid
Levulinic acid
Glycerol

transformed via aqueous-phase processing at mild temperatures.4


This approach allows us to carry out the required transformations in biomass platform molecules in the classical medium in
which they are obtained and in a controlled manner, avoiding
undesired reactivity leading to coke and loss of carbon which
typically occurs at elevated temperatures.
However, this route requires costly pretreatment and hydrolysis steps necessary to hydrolyze solid biomass to soluble
sugars, which can be further converted to platform molecules.
Consequently, the aqueous-phase approach is designed to
operate only with the carbohydrate fraction of biomass, and
this represents an important drawback compared to alternative
thermochemical approaches such as gasication and pyrolysis
which make use of the entire organic matter of this resource.
The catalytic strategies utilized to convert biomass derivatives and petroleum feeds into fuels and chemicals are, at the
same time, determined by their respective chemical nature.
Thus, the production of chemicals from petroleum typically
involves a previous functionalization step required to activate
the initial feedstock for subsequent upgrading processes
(Fig. 1), with the challenge being the addition of these functional groups in a selective fashion. This activation step is not
required for biomass platform molecules which show per se a rich
chemistry oering high exibility and numerous synthetic pathways for the production of a large variety of useful compounds.

Fig. 1

5268

In contrast, the low functionalization of petroleum feeds is


advantageous if the targeted product is a liquid hydrocarbon
fuel (e.g. gasoline, diesel and jet fuel). In this case, petroleum
derivatives can be used almost directly as feedstocks for the
production of these compounds after appropriate catalytic
processing (cracking, alkylation, isomerization) required to
control molecular weight and structure, viscosity and octane
or cetane number. On the other hand, the catalytic production
of liquid hydrocarbon fuels from biomass derivatives is a
complex process which involves deep chemical transformations
in a series of steps and requires of ecient strategies to
controllably remove functionalities and adjust the molecular
weight of the nal hydrocarbon fuel product.
The oxygen removal step is necessary to control the high
reactivity of intermediates and thus direct the synthesis to
targeted chemicals and fuels with high yields, as it will be
shown in some of the catalytic processes of the present work.
A variety of reactions including dehydration, hydrogenation,
decarbonylation/decarboxylation and CO hydrogenolysis
can be employed to reduce the oxygen content in biomass
platform molecules. Importantly, strategies to minimize the
utilization of fossil-fuel derived hydrogen (for example by
using a fraction of the feedstock as a source of hydrogen)
and to reduce CO2 emissions (by means of in situ sequestration
or capture technologies) during biomass deoxygenation steps

Comparative processing approaches of petroleum and biomass to chemicals and liquid hydrocarbon transportation fuels.

Chem. Soc. Rev., 2011, 40, 52665281

This journal is

The Royal Society of Chemistry 2011

Published on 29 June 2011. Downloaded by Universidad de Malaga on 28/10/2014 15:46:44.

View Article Online

are also important aspects to consider when designing biorenery processes. The deoxygenation process is often
combined with adjustment of the molecular weight via CC
coupling reactions (e.g. aldol-condensation, ketonization, oligomerization) of reactive intermediates and/or related transformations. Since biomass derivatives are typically molecules with
a maximum number of carbon atoms of six (derived from
glucose), these CC coupling steps acquire especial relevance
when the nal product is a larger hydrocarbon fuel, as those
currently used in diesel engines (C10C20) and jets (C9C16).

2. Catalysts for aqueous-phase conversion of


biomass platform molecules: new requirements
for new processing challenges
Modern industrial chemistry has been designed around petroleum,
and the processes and catalytic materials developed during the last
few decades are specically selected to operate with hydrocarbons.
Classical solid catalysts are materials with high thermal stability
and resistance to hydrophobic environments, in view of the
conditions required to process the relatively inert hydrocarbon
feedstocks. Inorganic solids such as alumina, silica and aluminasilicates meet most of those requirements and thus constitute the
basis of most catalysts utilised today in the petrochemical industry.
As indicated in the previous section, the special characteristics of biomass platform molecules suggest that these resources
would be more eciently processed at moderate temperatures
in the form of aqueous solutions, which places new requirements on the properties of the solid catalysts employed. Thus,
rather than thermally stable, these new catalysts for biomass
conversion should be stable under hydrothermal conditions
(moderate temperatures and high water concentrations).
Unfortunately, classical metal oxide supports are unstable at
these conditions and this issue is compounded when reactions
are carried out at basic or acid pH. Since traditional support
materials (e.g. alumina) strongly interact with water, they can
undergo irreversible structural changes and loss of textural
properties when exposed to hydrothermal conditions.
Consequently, alternative materials such as carbon, monoclinic zirconia, and rutile titania are preferred supports for the
aqueous-phase processing of biomass derivatives.5 Many of
the catalytic reactions involved in the transformation of platform molecules to fuels and chemicals are acid-catalysed
processes, as shown in subsequent sections. Since water typically poisons acid sites, the design of new water-tolerant solid
acids which keep their activity under hydrothermal conditions
is key to achieve ecient and cost competitive aqueous-phase
conversions of biomass derivatives. A variety of solid acids
including zeolites with high Si/Al ratio, heteropolyacids such
as H3PW12O40 and HNbMoO6, and oxides and phosphates of
niobium and zirconium are good examples on acidic materials
with high water compatibility. A comprehensive review with
the most relevant water-resistant solid acid catalysts can be
found elsewhere.6
Leaching of active sites is one of the most common issues in
biomass conversion processes.7 The strong chelating nature of
some components of biomass along with the use of acidic
conditions (normally required for hydrolysis or dehydration
This journal is

The Royal Society of Chemistry 2011

steps) can lead to severe dissolution of the catalytic active


phases by the reaction medium. Leaching causes irreversible
deactivation of catalysts and/or serious contamination of
reaction products. This issue is especially important if expensive noble metals are used as active phases or if the catalytic
process operates in a continuous form.
Carbon, apart from an excellent hydrothermal stability,
possesses high resistance to acidic and chelating media, and
thus it is an optimum material to prepare metallic catalysts for
aqueous-phase conversion of biomass derivatives. Additionally, it can be easily synthesized from inexpensive biomass
resources8,9 in the form of high surface area materials with
tuneable porosity and surface chemistry.9 In order to overcome leaching it is important to develop new techniques
for permanently anchor active sites to hydrothermal stable
supports as well as the design of new synthetic methodologies
to decrease the interaction of water with the active phases.
The variable and complex composition of biomass usually
involves a small fraction of inorganic components such as Cl,
F, P, S, Na and Mg. These impurities are dragged during the
hydrolysis step and, consequently, they are typically found in
water accompanying platform chemicals. These compounds,
even in trace amounts, can irreversibly poison catalytic sites.
For example, a few ppm of sulfur can rapidly deactivate a Ru
catalyst used for glucose hydrogenation,10 while the alkali
typically present in crude glycerol drastically reduces the
hydrogen yield of a platinum-based reforming catalyst.11
The design of new robust catalysts able to cope with these
impurities is thus imperative for the successful application of
aqueous-phase processing technologies. Since the nature of the
impurities can vary signicantly depending on the source and
quality of the raw material, it will be important to synthesise
catalysts with the broadest spectrum of tolerance possible in
order to ensure maximum applicability.
Because the chemical composition of the biomass feedstocks
is normally very dierent from that of the nal products,
multiple processing steps are also typically required in such
transformations, negatively aecting the economy of the process.
The development of multifunctional materials (with the ability
to catalyse several reactions in the same catalytic bed) can help
to reduce complexity of the transformations carried out in
bioreneries allowing the design of one-pot processes. The
use of a multifunctional Pd/Nb2O5 catalyst containing metal
and acid sites allows the transformation of aqueous solutions of
an important biomass derivative such as g-valerolactone into
valuable 5-nonanone in a single reactor.12 In this process, the
metal catalyses hydrogenation reactions while the acidic support
achieves simultaneous GVL ring-opening and ketonization of a
pentanoic acid intermediate. Relevant examples on the use of
multifunctional catalysts for platform chemicals valorisation
will be provided in subsequent sections.

3. Catalytic strategies to process biomass


feedstocks into fuels and chemicals
3.1.

Ethanol

The fermentation of edible biomass-derived aqueous sugars


to ethanol currently represents the main technology for the
Chem. Soc. Rev., 2011, 40, 52665281

5269

Published on 29 June 2011. Downloaded by Universidad de Malaga on 28/10/2014 15:46:44.

View Article Online

production of liquid biofuels in the world. The simple and wellknown fermentation process, along with a partial compatibility
with the existing infrastructure for gasoline, has allowed the
bioethanol industry to grow fast, and annual production has
almost doubled from 2007 (50 billion L) to 2010 (87 billion L).13
However, the utilization of ethanol in existing spark-ignition
engines is limited to low-concentration blends (typically 515%
by volume, denoted as E5 to E15). Ethanol-enriched mixtures
such as E85 require cars with specially designed engines, designated as exible-fuel vehicles (FFVs) which are common in a
few countries. For the US as the rst ethanol producer worldwide, this blend wall is currently generating issues to absorb the
ethanol production at the present time, and near future projections foresee an ever increasing expansion of the ethanol market
as a consequence of the implementation of technologies for
the conversion of more abundant lignocellulosic feedstocks.
Consequently, ethanol is likely to become a cheap and abundant
renewable resource in the near future and it was a particularly
timely platform molecule choice in Bozells revisited list.3
Ethanol has an enormous potential as a chemical feedstock
(Fig. 2). It can serve as a source of renewable hydrogen
(currently obtained from fossil fuels) via steam reforming
processes.14 This route, widely investigated in past years,
involves the gasication of aqueous solutions of ethanol at
high temperatures (typically 600800 1C) and atmospheric
pressure using Ni, Co and noble metals (e.g. Pt, Pd, and Rh)
supported on stable oxides. Hydrogen is co-produced along
with CO2 with a maximum yield of 6 moles of gas per mol of
ethanol fed. The main challenges of this technology originate
from the high temperatures required to favour the endothermic reforming. Thus, catalysts stability (fast deactivations
by coke deposition are frequently found in the literature) and
control over formation of by-products (typically CH4 and CO)
are negatively aected by harsh temperature conditions.

Ethanol is readily dehydrated, even under aqueous environments,15 over acidic catalysts at moderate temperatures and
with almost quantitative yields, serving as a renewable source
of ethylene, the most produced organic compound worldwide
(annual production 120 million tons) and one of the seven
primary building blocks of the petrochemical industry.
Current oil prices and the rapid development of fermentation
technologies have increased the attractiveness of the bioethylene route in comparison with the classical steam-cracking
petrochemical process. The production of greener polyethylenederived plastics, demanded by chemical users and costumers, is
another incentive for the development of this route. Recently,
Brazilian Braskem A.S. inaugurated the rst commercial-scale
ethanol-to-ethylene plant which will produce 200 000 tons
per year of green polyethylene from sugar cane ethanol.16
Renewable ethanol can also be transformed into of butadiene,
another important petrochemical building block utilized as a
raw chemical in the industry of synthetic rubber.17 The technology for the conversion of ethanol into butadiene is not new
(the process has been utilized since World War II to produce
rubber at small scale), but it was traditionally replaced by the
more economical petroleum steam cracking route. The costeective production of primary olen building blocks such as
ethylene and butadiene converts ethanol in an interesting bridge
molecule between bioreneries and current petrochemical
industry, facilitating a less drastic transition between both
concepts. Propylene, the second most important starting product
in the petrochemical industry after ethylene, can be also
produced from biomass-derived ethanol. The process involves
dehydration to ethylene, partial dimerization of the latter to butene
and subsequent metathesis of both C2 and C4 olens to yield
propylene.18 Removal of nitrogen and sulfur containing compounds below the ppm level from the ethylene stream is crucial
since metathesis catalysts are highly-sensitive to these impurities.

Fig. 2 Production of fuels and high added value chemicals from ethanol.

5270

Chem. Soc. Rev., 2011, 40, 52665281

This journal is

The Royal Society of Chemistry 2011

Published on 29 June 2011. Downloaded by Universidad de Malaga on 28/10/2014 15:46:44.

View Article Online

The potential of ethanol as a bridge molecule could be boosted if


eective routes to other more complex petrochemical building
blocks (e.g. benzene, toluene and xylenes) are developed, for
example, by controlled oligomerization of ethylene.19
As represented in Fig. 2, important C2 commodity chemicals
can be also derived from ethanol. Dehydrogenation to acetaldehyde is performed with 100% selectivity using inexpensive
Cu catalysts at mild temperatures and ambient pressure20 in
an attractive process that would allow simultaneous production
of renewable hydrogen and acetaldehyde (a valuable chemical
produced on large scale) in a simple and clean one-step reaction.
Christensen et al. reported an aqueous aerobic oxidation
methodology to convert dilute ethanol into acetic acid using
supported Au catalysts at moderate temperatures (150 1C) and
pressures.21 Acetic acid (annual production 6.5 million of
tonnes) is an important chemical with multiple applications in
the elds of polymers and food industry, among others. Most of
the world production is currently obtained by carbonylation
of petroleum-derived methanol using homogenous Ir and Ru
metal complexes. Ethanol can thus nd a market in the
production of acetic acid by oering a green, simple and costcompetitive alternative to the conventional fossil fuel route.
Furthermore, the selectivity can be switched to produce ethyl
acetate (an important organic solvent manufactured on large
scale worldwide), if reaction conditions are adjusted (e.g.
increasing ethanol concentration).22
3.2.

Furans

3.2.1. Hydroxymethylfurfural (HMF). Biomass-derived


hexoses, obtained from cellulose, hemicellulose or starches,
can be dehydrated to form furan compounds such as HMF.
Its formation implies the removal of 50% of the oxygen
(via water elimination) from the initial sugar feedstock in a
process typically carried out in aqueous mineral acids such as
HCl or H2SO4. However, the use of homogenous acids has a
number of associated drawbacks, namely diculties to recover
the acid upon reaction completion and the requisite to employ
expensive reactors resistant to corrosion. Current research is
more focused on the utilization of solid acids which: (i) are
easily separable and recyclable from the reaction medium;
(ii) allow higher reaction temperatures without damaging the
reactor, thereby shortening reaction time and decreasing the
amount of decomposition by-products; and (iii) can be synthesized with a range of tuneable surface acidities and porosity
properties to improve HMF selectivity. In this sense, a variety
of water-tolerant solid acids including H-form zeolites, ionexchange resins, vanadyl and niobium phosphates, and ZrO2
have been tested in the lab-scale dehydration of sugars to
HMF.23
Cheap and abundantly available glucose is the desirable
starting material for HMF production, although yields are
typically low for this feedstock. To overcome this limitation,
current technologies employ an additional glucose isomerization step to fructose, which is more easily converted into the
furanic intermediate. In any case, the main drawback involved
in the dehydration of hexoses to HMF is the high number of
unwanted side reactions, which need to be controlled in order
to maximize the nal HMF yield. The use of biphasic reactors
This journal is

The Royal Society of Chemistry 2011

to continuously extract the HMF into an organic phase thus


preventing the furanic intermediate from overreacting in the
aqueous phase,24 and the utilization of ionic liquids as reaction
media25 are some of the most promising approaches to improve
HMF yields.
HMF is currently utilised as feedstock for the production of
industrial solvents and it has a tremendous potential as
substitute of petroleum-based building blocks in the synthesis
of large-scale polymers including polyesters, polyamides and
polyurethane. HMF could also nd use in the transportation
sector as renewable additive in conventional hydrocarbon
fuels, in a similar way to that of ethanol (Section 3.1).
However, its excessive oxygen content which leads to a high
boiling point, high solubility in water, low energy-density and
high-reactivity characteristics makes this molecule unattractive as liquid fuel per se.
One eective approach to achieve HMF deoxygenation
involves hydrogenation of the carbonyl group and subsequent
CO hydrogenolysis of the hydroxyl intermediate dihydroxymethyl furan (DHMF)26 (Fig. 3). The resulting product,
dimethylfuran (DMF), is hydrophobic and possesses excellent
energy-density and boiling point characteristics to be used as
transportation liquid. Since this reaction is carried out under
hydrogen pressure, the challenge is to achieve the removal of
the hydroxyl groups in DHMF while preserving the CQC
bonds in the furanic ring (which otherwise lead to unnecessary
hydrogen consumption). The use of a copper-based catalyst
allows this transformation of HMF into DMF to selectively
take place in a single step. A slow deactivation with time on
stream is reported, although activity can be fully recovered by
owing H2 at reaction temperature. The main challenge for
the large-scale implementation of this technology is the cost of the
process which is negatively aected by the complexity of the
HMF production/purication steps. A recent breakthrough,
which allows us to dramatically reduce complexity in the conversion of fructose to DMF, is reported by Thananatthanachon
and Rauchfuss.27 In this work, formic acid, a by-product from
biomass degradation processes, is utilised as: (i) a homogeneous
acid to achieve fructose dehydration to HMF; (ii) a source of
hydrogen for HMF reduction to DHMF; and (iii) a catalyst to
achieve deoxygenation of DHMF to DMF. This promising
technology achieves one-pot conversion of fructose into DMF
with acceptable yields (51%).
One alternative strategy to upgrade HMF involves the
utilization of the carbonyl group as a reactive centre for the
production of larger hydrocarbon fuels. Linear alkanes with
molecular weights appropriate for diesel and jet fuel applications (C9C15) can be derived from HMF by a cascade process
involving dehydration, hydrogenation and aldol-condensation
reactions28 (Fig. 3). This technology requires an external
carbonyl-containing molecule (typically acetone) to initiate
the condensation step, which is usually carried out in polar
solvents such as water using homogeneous basic catalysts.
The generated aldol-adduct can additionally undergo a second
condensation with the original furanic molecule to increase even
more the molecular weight of the nal product. Intermediate
aldol-adducts are subsequently converted into water-soluble
compounds aiming to stabilise them. Finally, hydrogenated
aldol-adducts are converted into liquid alkanes via aqueous-phase
Chem. Soc. Rev., 2011, 40, 52665281

5271

Published on 29 June 2011. Downloaded by Universidad de Malaga on 28/10/2014 15:46:44.

View Article Online

Fig. 3 HMF as a platform molecule for the production of high-added value chemicals and biofuels.

dehydration/hydrogenation (APD/H) reactions over a


bifunctional metal-acid catalyst. The need for an external
carbonyl-containing agent (typically derived from petroleum)
and the large number of steps required to achieve complete
deoxygenation and molecular weight adjustment of the
nal hydrocarbon product are the main drawbacks of this
technology.
HMF, along with ethanol, will play a key role in the bioplastics and bio-polymers eld. As indicated in Section 3.1,
sugar cane-derived ethanol is presently utilised for the production of polyethylene (PE), the most widely used plastic in the
world with multiple applications in packaging. Bio-ethylene
can alternatively be utilised to produce ethylene glycol
(1,2-ethanediol) via hydration of ethylene oxide. Subsequent
esterication of ethylene glycol with terephthalic acid (benzene1,4-dicarboxilic acid) yields polyethylene terephthalate (PET),
another important polymer massively employed in the manufacture of synthetic bers and plastic bottles. HMF can nd
market in the production of bio-PET by oering a green
substitutive for petroleum-derived terephthalic acid, which is
employed in high amounts (e.g. 70% by weight) in the PET
industrial process. This substitutive is 2,5-furandicarboxylic
acid (FDCA) which can be produced by aerobic oxidation of
HMF (Fig. 3) and possesses chemical structure and properties
similar to terephthalic acid. The catalytic transformation of
HMF into FDCA involves reaction under pressurized air/oxygen
at mild temperature conditions (100150 1C) and controlled
5272

Chem. Soc. Rev., 2011, 40, 52665281

alkaline pH. Platinum29 and gold-based30 catalysts have been


reported to achieve nearly quantitative yields of FDCA in water.
The one-pot fructose dehydration and HMF oxidation steps
have also reported to be feasible using a Co acetylacetonate/SiO2
bifunctional (acid and redox) catalyst, although conversion was
limited to 72%.31 The conversion of HMF to FDCA is a very
favourable reaction which can even be carried out at ambient
temperature32 and without utilization of external homogenous
base.33 However, the success of this technology depends, as in the
case of DMF, on the design of ecient routes for HMF
production. Furthermore, the eect of the new FDCA monomer
on the properties of bio-PET is relatively unknown.
3.2.2. Furfural. Furfural (furan-2-carbaldehyde), with an
annual production volume of more than 200 000 tons, is one of
the most common industrial chemicals derived from lignocellulose and a key intermediate for the complete utilization
of hemicellulose-rich biomass feedstocks such as bagasse, corncobs and stalks, switchgrass and hardwood. The importance
and potential of furfural as renewable feedstock is signicant as
currently there is no synthetic route for the direct production of
this molecule. Furfural is the dehydration product of xylose,
a C5 sugar typically found in hemicellulose.
The commercial process for the production of furfural is
based on the use of mineral acids (H2SO4) at moderate
temperatures (o200 1C). The homogenous acids (in the form
of dilute aqueous solutions) achieve a one-step hemicellulose
This journal is

The Royal Society of Chemistry 2011

Published on 29 June 2011. Downloaded by Universidad de Malaga on 28/10/2014 15:46:44.

View Article Online

deconstruction and xylose dehydration in the same reactor.


However, as indicated in the previous section, the use of
mineral acids is troublesome, and furfural yield is rarely above
50 mol%. New technologies based on the use of organic
solvents to continuously extract furfural from the aqueous
phase once produced and thus avoid losses by overreaction,34
ionic liquids to assist in hemicellulose depolymerization,35 and
easily-separable acidic solids including zeolites, hydrotalcites,
heteropolyacids and acidic resins to achieve xylose dehydration36
are currently being explored to improve furfural yields and to
design a cleaner and more environmental friendly industrial
process.
Fig. 4 shows the high platform potential of furfural for the
simultaneous production of fuels and important chemicals in
bioreneries. Hydrogenation is the most common route to
upgrade furfural, from which a number of important products
can be synthesized in high yields depending on the extent
of the reduction step. Thus, reduction at mild conditions
(e.g. 120 1C, atmospheric pressure) using Cu-based catalysts
yields furfuryl alcohol,37 the most important derivative of
furfural with high demand in the manufacture of foundry
resins. Furfuryl alcohol can additionally serve as a feedstock
for the production of levulinic acid (4-oxopentanoic acid), an
important biomass platform molecule which will be analyzed
in Section 3.3.3. The process involves hydrolysis of furfuryl
alcohol at mild temperatures under aqueous acids which
typically yields 80% of levulinic acid.38
Harsher hydrogenation conditions including higher pressures
(20 bar) and/or the use of supported noble-metal catalysts lead
to complete hydrogenation of the furanic ring to produce methyl
tetrahydrofuran (MTHF) (Fig. 4). MTHF is a hydrophobic

molecule which, unlike ethanol, can be blended with gasoline up


to 60% (v/v) without adverse eects on engine performances or
gas mileage. Furthermore, MTHF is an excellent substitutive for
dichloromethane (DCM), an usual solvent in pharmaceutical
production and a suspected carcinogen. MTHF presents
superior extraction characteristics for polar compounds, and
its higher boiling point (80 1C) allows easier condensation/
recycling compared to more volatile DCM. The use of Pd-based
catalysts facilitates decarbonylation versus hydrogenation processes, aording furan which is readily reduced to tetrahydrofuran (THF).39 THF is a common organic solvent and precursor
of polymers with an annual production in excess of 200 000 tons.
Furfural can also serve as a good source of carboxylic acids
via oxidation. Thus, furoic acid (with market in the pharmaceutical and the agrochemical elds for the production of
drugs and insecticides) and maleic acid (an important raw
chemical not found in nature used in the production of alkyl
resins with polyols such as glycerol) can be synthesized with
acceptable yields by using the oxidative approach.40
In a route that imitates that of HMF (Section 3.2.1),
furfural can be transformed into C8C13 diesel and jet fuel
hydrocarbon components by means of dehydration, hydrogenation and aldol-condensation cascade processes28 (Fig. 4).
Recent improvements in the process have increased the attractiveness of this route, with projections to produce diesel and jet
fuel range alkanes from hardwood-derived hemicellulose at a
price of $0.531.16 L.41 However, the price of raw materials,
the organic-to-aqueous mass ratio in the biphasic dehydration,
and xylose yield after hemicellulose depolymerization are
sensitive parameters which need to be controlled to ensure
economic feasibility.

Fig. 4 Chemical transformations of furfural as a platform molecule into liquid biofuels and chemicals.

This journal is

The Royal Society of Chemistry 2011

Chem. Soc. Rev., 2011, 40, 52665281

5273

View Article Online

Published on 29 June 2011. Downloaded by Universidad de Malaga on 28/10/2014 15:46:44.

3.3.

Organic acids

3.3.1. Lactic acid. Lactic acid (2-hydroxypropanoic acid,


LA) is the most widely occurring carboxylic acid in nature.
Even though lactic acid can be synthesized by hydrolysis of
petroleum-derived lactonitrile, the large majority of the annual
production of this feedstock (120 000 tons per year) is obtained
by bacterial fermentation of both starchy and lignocellulosic
biomass. Lactic acid has been traditionally used as an additive
in the food industry although new applications in the eld
of chemicals and polymers production have been recently
developed for this platform molecule.42
The bifunctional character of lactic acid (e.g. OH and
COOH groups) allows a variety of transformations to useful products (currently obtained from petroleum) including
acrylic acid (an important raw material for the production of
plastics, adhesives and paints) via dehydration, propanoic
acid (via dehydration/hydrogenation), acetaldehyde (via
decarbonylation/decarboxylation), 2,3-pentanedione (via condensation) and polylactic acid (PLA), a new biodegradable
polymer with excellent properties for the fabrication of plastics
and related materials (Fig. 5). These new applications, the
improvement of existing fermentation and separation technologies, and the development of non-biological routes for the
conversion of aqueous sugars into lactic acid based on easilyseparable and recyclable solid zeolites43 have considerably

increased the platform potential of this molecule in recent


years.
A partial oxygen removal approach has been recently
applied to catalytically process lactic acid into C4C5 alcohols
suitable as high energy-density liquid fuels for the transportation sector44 (Fig. 6). LA is deoxygenated in a rst step to
generate two reactive intermediates, namely propanoic
acid and acetaldehyde, by means of CO (dehydration followed
by hydrogenation) and CC cleavage (decarbonylation/
decarboxylation) processes, respectively. Once the two intermediates have been formed, they can be upgraded to larger
compounds by means of CC coupling reactions. Acetaldehyde
can thus undergo self-coupling by aldol-condensation to
generate butanal (after hydrogenation of the corresponding
C4 unsaturated aldol-adduct), whereas propanoic acid can be
catalytically converted into 3-pentanone via ketonization. Both
butanal and 3-pentanone can then be readily hydrogenated
to their corresponding alcohols. Remarkably, the series of
reactions leading from lactic acid to butanal and 3-pentanone
can be carried out in a single reactor by employing multifunctional Pt/Nb2O5 catalyst in which Pt achieves hydrogenation and acidic niobia catalyses dehydration, decarboxylation,
and CC coupling reactions. Concentrated aqueous solutions of
lactic acid (60 wt%) were eciently processed over Pt/Nb2O5
with good stability for more than two days on stream.
The development of cost-eective technologies for lactic acid

Fig. 5 High-added value chemicals derived from various transformations of lactic acid.

5274

Chem. Soc. Rev., 2011, 40, 52665281

This journal is

The Royal Society of Chemistry 2011

Published on 29 June 2011. Downloaded by Universidad de Malaga on 28/10/2014 15:46:44.

View Article Online

Fig. 6

Processing of lactic acid into alcohols suitable as liquid biofuels.

production will determine the success of this route in near


future.
3.3.2. Succinic acid. Succinic acid (SA) is a 1,4-dicarboxylic
acid, which repeating in both the original USDOE and
Bozells list, has been demonstrated to be one of the key
platform molecules to be transformed into a variety of useful
chemicals as a replacement of maleic anhydride. From esters
to amides through pyrrolidones, alcohols and/or biopolymers,
the rich chemistry of the two carboxylic groups within the
molecule (Fig. 7) as well as its partial solubility in water are
two key assets for relevance from the list of building blocks.
For further reading on transformations of succinic acid
into useful chemicals, readers are kindly referred to detailed
literature reviews from Cukalovic and Stevens45 and
Delhomme et al.46
Traditionally, SA is produced from petroleum via oxidation
of n-butane but it has been recently obtained from various
biomass feedstocks including wheat, corn waste and others.47
DSM and Roquette started up a joint venture (Reverdia v.o.f.)
to produce bio-based succinic acid building up from previous
experience in the SA demonstration plant in Lestrem (France)
built in 2009.
Webb et al. have also shown that SA can be either isolated
from fermentation aqueous broths or obtained at ca. 4060 g L 1
concentration in the broth (46 wt% SA) at the laboratory
scale.48 More recent studies have shown interesting concentrations of SA (up to 40 g L 1) can be produced from a seawater
fermentation broth using low cost biomass-derived raw materials
(e.g. waste from cereals) which strictly speaking will tick all
boxes of the biorenery paradigm (renewable, cheap and widely
available water source combined with a low cost renewable
feedstock).49
The SA thus obtained can be feasibly subjected to a series
of transformations summarised in Fig. 7. Esterications,
This journal is

The Royal Society of Chemistry 2011

amidation and hydrogenations all performed in aqueous media


are proven transformation which occurred in high yields
and selectivities using Starbons materials as water tolerant
catalysts.50 Of particular interest are aqueous esterications and
amidation reactions. Both processes are reversible equilibria in
which water is critically involved in such a way that it was
traditionally required to be removed from the system (e.g. via
dean stark or capturing by molecular sieves and/or desiccants)
to obtain good yields to products. The water-compatibility
of these systems allows the possibility to work with dilute
aqueous mixtures in which less water soluble formed products
(e.g. esters, amides, lactones) are easily separated out from the
original mixtures.
The obtained high added value chemicals including diethyl
succinate, g-butyrolactone, THF and succinamides have
important applications as solvents, additives in cosmetics
and fragrances, biopolymers, etc. Succinic acid hydrogenation
has also been reported to be promoted by supported noble
metals (e.g. Pd, Ru) for the selective production of g-butyrolactone (GBL).51 A Pd on mesoporous alumina xerogel catalyst could reach a 79% conversion of succinic acid for a 51%
yield to GBL in 4 h of reaction at 60 bar of hydrogen.51a
Comparatively, a lower yield to GBL (30%) at comparative
conversion of SA was reported after 24 h reaction using Pd
nanoparticles supported on Starbons under mild reaction
conditions (aqueous media, 10 bar hydrogen pressure,
80100 1C).51b Ru/C catalysts doped with others metals
(e.g. Co) have also been reported to give GBL in high yields
as originally reported by Deshpande et al.51c However, these
materials only provided relatively good activities under forced
conditions of temperatures.
3.3.3. Levulinic acid. The treatment of biomass-derived
hexoses at moderate temperatures under aqueous acid
conditions leads to the production of HMF (Section 3.2.1).
Chem. Soc. Rev., 2011, 40, 52665281

5275

Published on 29 June 2011. Downloaded by Universidad de Malaga on 28/10/2014 15:46:44.

View Article Online

Fig. 7 Production of high-added value chemicals and precursors to biofuels from succinic acid.

This compound is however highly reactive and readily decomposes into an equimolar mixture of formic and levulinic acids
(LVA) if overreaction in the aqueous phase is allowed. LVA
can thus be obtained in high yields via acid hydrolysis of C6
sugars using technologies such as the Bione process,52
which achieves industrial production of this compound from
inexpensive lignocellulosic wastes including paper mill sludge,
urban waste paper, and agricultural residues at a low cost.
Because of the co-production of formic acid during the
dehydration process, LVA is generated with a theoretical
maximum yield of 64.5% by weight. However, this maximum
yield is unlikely to be reached since approximately 1/3 of the
carbon in the initial sugar feedstock is typically lost as
unwanted black insoluble materials, the so-called humins.
These intractable solids are normally burnt to generate heat
and electricity for the process.
LVA could be also potentially obtained from hemicellulosederived pentoses via dehydration to furfural and subsequent
hydrogenation to furfuryl alcohol (Fig. 4). This route would
benet from an easier deconstruction of amorphous hemicellulose compared to highly-recalcitrant crystalline cellulose.
Additionally, although the production of LVA from hemicellulose requires an additional hydrogenation step which is
5276

Chem. Soc. Rev., 2011, 40, 52665281

not required with cellulose, this reaction is readily achieved


with yields higher than 95% in a well established industrial
process.37 On the other hand, hemicellulose represents a minor
fraction of lignocellulose (2035%) in comparison with cellulose (4050%) and, unlike cellulose, it possesses a heterogeneous composition (involving dierent sugars and acetyl
groups) that complicates the hydrolysis step and subsequent
furfural purication.
The development of new aqueous-phase technologies for the
catalytic conversion of LVA into liquid transportation fuels
has boosted its platform potential in recent years (Fig. 8).
The key intermediate which allows ecient transformation of
LVA into fuels is g-valerolactone (GVL). GVL is a stable
and water-soluble interesting biomass-derivative that has been
proposed to be a potential gasoline additive as well as a
precursor of polymers and ne chemicals.53 GVL can be
produced with almost quantitative yields via hydrogenation
of aqueous LVA at low temperatures (o200 1C) utilising
non-acidic catalysts (e.g. Ru/C) through the intermediate
4-hydroxypentanoic acid.12 The utilization of higher temperatures and/or acidic catalysts generates increasing quantities
of angelica lactone (AL), a well known coke precursor.
Interestingly, as in the case of the conversion of HMF to
This journal is

The Royal Society of Chemistry 2011

Published on 29 June 2011. Downloaded by Universidad de Malaga on 28/10/2014 15:46:44.

View Article Online

Fig. 8 Levulinic acid as a platform molecule for the production of liquid transportation biofuels.

DMF (Section 3.2.1), the formic acid co-produced in the sugar


dehydration process can serve as a hydrogen source for LVA
reduction to GVL, thereby avoiding utilization of external
fossil-fuel derived hydrogen.54
As summarized in Fig. 8, GVL possesses a high versatility to
synthesize liquid transportation fuels of diverse classes. It can
be readily converted to MTHF via hydrogenation with the
intermediate formation of 1,4-pentanediol.55 Alternatively,
aqueous GVL can be transformed into hydrophobic pentanoic
acid (PA) in high yields using a bifunctional Pd/Nb2O5
catalyst via ring-opening (on acid sites) and hydrogenation
(on metal sites) reactions at moderate temperatures and
pressures.12 This route has been recently exploited by Lange
et al.56 to produce valeric biofuels (i.e. alkyl valerates), a
new family of lignocellulosic biofuels with excellent energy
density, polarity and volatility-ignition properties to be used in
conventional engines without any modication. Remarkably,
the volatility-ignition properties of such valerates can be
adapted to t in both gasoline and diesel engines by selecting
adequate alkyl chain lengths. Pentanoic acid can alternatively
be converted into 5-nonanone in high yields (90%) via ketonization reactions.12 The C9 ketone serves then as a platform
molecule for the production of hydrocarbon fuels with molecular weight and structures adequate for gasoline, diesel and
jet fuels applications.
Recently, a promising route to upgrade aqueous solutions
of GVL into jet fuels through the formation of C4 alkenes has
been developed by Bond et al.57 This dual-reactor process
achieves GVL decarboxylation to butenes at 36 bar over
inexpensive silicaalumina. Butenes are subsequently oligomerised using acidic Amberlyst as catalyst, yielding a distribution of alkenes centred at C12 which is potentially suitable for
jet fuels applications. The process is simple (only two reactors
This journal is

The Royal Society of Chemistry 2011

working in series), clean (CO2 can be eciently sequestrated at


the system pressure) and requires minimum utilization of
external hydrogen. However, liquid water must be removed
from the gas stream in a separator to achieve eective oligomerization of butene in the second reactor.
3.4.

Polyols and sugars

3.4.1. Glycerol. Glycerol (1,2,3-propanetriol) is a high


boiling point (290 1C), water-soluble biomass-derived polyol
whose importance as a renewable platform molecule has
dramatically increased in recent years.58 Glycerol is daily
produced in large quantities in biodiesel facilities (100 kg of
glycerol per ton of biodiesel) as byproduct of the transesterication of vegetable oils and animal fats with methanol.
The rapid growth of the biodiesel industry (world production
reached 16 billion L in 2009, and it is expected to increase to
45 billions L by 202058b) is generating a surplus of crude
glycerol in the market which resulted in a signicant drop
of prices for this feedstock (from $0.55 per kg in 2004 to
$0.055 kg in 2006). This has not only pressurised producers of
glycerol but most importantly biodiesel producers who cannot
sell crude glycerol at a price high enough to compensate the
high production cost which in turn could make biodiesel less
economically competitive as compared to petrol and diesel.
However, this uctuation in crude glycerol prices can be
benecial for its application in other markets. It is clear that
three of the main and traditional applications of glycerol
(food, health care and pharmaceuticals which constitute more
than 50% of todays glycerol applications) are signicantly
restricted for crude glycerol unless a previous purication step
is carried out. Nevertheless, there are other markets for which
crude glycerol may be very suitable including the production
Chem. Soc. Rev., 2011, 40, 52665281

5277

Published on 29 June 2011. Downloaded by Universidad de Malaga on 28/10/2014 15:46:44.

View Article Online

of polyols, renewable plastics and resins and detergents, for


which the quality and purity requirements are generally lower.58
In any case, it is worth pointing out at this stage that world
trading and price volatility of glycerol can signicantly aect its
value for its transformation in high added value products and
a biodiesel crisis (mainly responsible of the excess of crude
glycerol) can increase its prices and therefore drive back the uses
of glycerol to traditional uses and/or direct selling.
In spite of these considerations, it is evident that the
expansion of the biodiesel industry has lead to abundant and
cheap glycerol for which novel technologies for the large-scale
processing of this resource into valuable products have been
increasingly investigated in recent years and considered to be
of primary importance.
Glycerol exhibits a versatile and high chemical reactivity
due to its three hydroxyl groups. Some of the most important
and novel routes to upgrade glycerol into fuels and useful
chemicals are depicted in Fig. 9. Similarly to ethanol (Section
3.1), aqueous glycerol can be gasied to CO, H2 and CO2 via
reforming processes using metal-based catalysts.59 The aqueousphase reforming of glycerol oers a number of technical
advantages as compared to ethanol. Firstly, the higher reactivity
of glycerol in comparison with ethanol allows the reforming
process to be carried out at milder temperatures (typically
200280 1C). Unwanted reactions are kinetically restrained
under these conditions, which facilitates an optimal control of
the gas-phase composition and improves catalysts lifespan.
Secondly, the process is more versatile and reaction conditions
and catalytic materials can be carefully selected to transform
concentrated aqueous solutions of glycerol (as those produced
in biodiesel facilities) into either syngas (a valuable mixture of
CO and H2) for Fischer-Tropsch (F-T) or methanol syntheses,
or, alternatively, H2-enriched streams by coupling APR with
water gas-shift (WGS) processes.

Important commodity chemicals such as C3 diols can be


also produced from glycerol by means of CO hydrogenolysis
reactions (Fig. 9). These processes, typically involving
dehydration and subsequent hydrogenation of the unsaturated
intermediates (e.g. acetol and 3-hydroxypropanal), currently
represent one of the most promising routes to valorise crude
glycerol. Propylene glycol (1,2-propanediol), the main product
obtained by this route, nds market in dierent sectors as
polyester resins, antifreeze, pharmaceutical, and paints,
among others, with a global demand of 1.6 billions of tons
in 2007. The commercial process uses a copper-chromite
catalyst at mild temperatures and hydrogen pressures (e.g.
200 1C and 10 bar) in a two step route involving dehydration
of glycerol to acetol and subsequent hydrogenation of the
carbonyl intermediate to glycol.58a The use of two reactors is
justied by an excellent nal glycol yield (95%) and the
possibility of isolating the valuable intermediate acetol. Again,
the use of mild conditions is key to selectively stop the
synthesis in propylene glycol, which typically overreacts to
lower value-added alcohols (e.g. 1- and 2-propanol) at harsher
conditions of pressure and temperature. Higher temperatures
and pressures along with the use of noble metals achieve,
however, CC hydrogenolysis to yield ethylene glycol,60 a
classical antifreeze liquid and precursor of polymers (Section
3.2.1) utilised worldwide.
Glycerol can also be dehydrated in the aqueous phase to
acrolein, an important chemical which serves as an intermediate in the production of acrylic acid, under high acidic
conditions (e.g., mineral acids such as H2SO4 or strong solid
acids including ZrO2WO3 or zeolites).58a,61 Acrylic acid
(annual production 1.2 millions tons) is readily polymerized
to acrylic resins which are widely used as superabsorber
polymers. The acidic conditions required to remove two
of the hydroxyl groups in glycerol also favour unwanted

Fig. 9 Main routes for the aqueous-phase transformation of glycerol into fuels and chemicals.

5278

Chem. Soc. Rev., 2011, 40, 52665281

This journal is

The Royal Society of Chemistry 2011

Published on 29 June 2011. Downloaded by Universidad de Malaga on 28/10/2014 15:46:44.

View Article Online

reactions such as polymerization and cracking, which lead


to catalyst deactivation by coke deposition. Consequently,
acrolein yields are usually moderate (acetaldehyde, propanal
and acetone are also produced in signicant amounts), and
catalysts regeneration (by burning-o the coke under oxidant
atmosphere) frequently needed.
Glyceraldehyde and dihydroxyacetone, two intermediates
that can be readily formed by gentle oxidation of aqueous
glycerol under air, can be transformed into lactic acid via
isomerization reactions using zeolite materials.62 This work
is relevant, as it oers a simple, clean and non-enzymatic
alternative to classical fermentation routes for the large scale
production of lactic acid.
3.4.2. Other polyols and sugars. Biomass-derived sugars
are the primary feedstocks for the production of important
platform chemicals. C5 and C6 carbohydrates are thus

converted, through both chemical and biological processes,


into stable and simpler intermediates which are subsequently
more easily handled and processed to the nal product.
Alternatively, sugars and primary derivatives such as polyols
(obtained in quantitative yields by hydrogenation of the
corresponding carbohydrates) can be directly processed
in the aqueous phase to valuable nal products thereby
eliminating intermediate steps and reducing complexity during
biomass conversion processes. For example, sugars can be
reformed in the aqueous phase to H2 using platinum-based
catalysts63 without previously converting them into ethanol
(see Section 3.1), and polyols such as xylitol and sorbitol can
undergo CC hydrogenolysis to yield valuable C2 and C3
glycols in a process that resembles that described in the
previous section for glycerol upgrading.64
Lactic acid, classically manufactured by bacterial fermentation, can be also synthesized directly from sugars (e.g., fructose,

Fig. 10 Scheme of the process for the catalytic conversion of sugars and polyols into liquid hydrocarbon fuels over PtRe/C.

This journal is

The Royal Society of Chemistry 2011

Chem. Soc. Rev., 2011, 40, 52665281

5279

Published on 29 June 2011. Downloaded by Universidad de Malaga on 28/10/2014 15:46:44.

View Article Online

glucose and sucrose) utilising solid Sn-beta zeolites43 with


glyceraldehyde and dihydroxyacetone (as in the case of
glycerol) being the key intermediates in the process. This
promising technology achieves improvements over fermentation by facilitating separation/collection of the lactic acid
product and recycling of the solid catalyst.
Biomass-derived sugars and polyols can serve as feedstocks
for the production of liquid hydrocarbon fuels. Unlike the
prevalent sugars-to-ethanol approach, this route allows the
production of hydrophobic, high-energy density fuels fully
compatible with current transportation infrastructure.
However, this paradigmatic transformation requires deep
oxygen removal from the feedstock (Fig. 1). An attractive
strategy to achieve deoxygenation of aqueous sugars and
polyols involves repetitive cycles of APD/H over water-stable
Pt/NbPO4 bifunctional metalacid catalysts.28 Since APD/H
only involves dehydration and hydrogenation reactions, it is
limited to the production of alkanes with the same carbon
atoms of the initial feedstock. This limitation was recently
overcome by Kunkes et al.65 by developing a two-step cascade
process that combines oxygen removal and CC coupling
upgrading (Fig. 10). In a rst step, sugars and polyols are
partially deoxygenated (up to 80%) by CO hydrogenolysis
processes on a PtRe/C catalyst at temperatures around
200 1C to yield a mixture of monofunctional hydrocarbons
in the C4C6 range (including acids, alcohols, ketones and
heterocycles) which separate out from water upon formation.
Remarkably, the hydrogen required to accomplish deoxygenation is internally supplied by aqueous-phase reforming
of a fraction of the feed over the multifunctional PtRe/C.
In a subsequent cascade process, the organic stream of monofunctionals is upgraded to targeted hydrocarbons through
dierent CC coupling reactions such as aldol-condensation
and ketonization. The high cost of the PtRe (10 wt%)/C
catalyst represents an important drawback for the implementation of this technology, and additional studies are
required to better understand the eect of catalysts formulation and feedstock type on the composition of the organic
stream of monofunctionals.

4. Future prospects and outlook


Biomass valorisation is a relevant area of research which will
continue attracting a critical mass of scientists in the near
future due to the remarkable potential of biomass as an
alternative source of chemicals, fuels and related commodities
as compared to petrol. However, working with complex
biomass feedstocks (e.g. lignin, waste raw materials) has been
proved to be rather dicult and advances in the eld are
mostly slow and under development. In view of these premises,
platform molecules can serve as an interesting starting point to
realise the biomass valorisation concept and ultimately the
biorenery paradigm.
A deeper and more profound understanding of processes,
mechanisms and transformation pathways of such building
blocks will be highly useful to set the basis of the behaviour of
more complex systems (e.g. lignin, lignocellulosics) under
similar conditions. The aqueous phase processing of platform
molecules is however not a simple task. Ecient, reusable and
5280

Chem. Soc. Rev., 2011, 40, 52665281

water-tolerant systems are required to promote these chemistries which become particularly challenging in dilute aqueous
broths containing a range of impurities apart from the target
compound. Furthermore, designer and tuneable materials and/
or catalysts are often required and these need to be carefully
designed and optimised for the chosen transformation.
Highlighting a series of general examples of these processes
and/or catalysts has been the subject of this work, in which
we hope the readers have been given a broad and dynamic
overview of the dierent routes to high added value chemicals
and fuels from the key building blocks possessing greater
potential for future development.
In any case and despite the important challenges faced
by this eld, the future of chemistry in these relevant tasks
(though still being far to be optimum, feasible and fully
understandable) is remarkably bright. Recent reports from
various authors on the production of renewable chemical
commodity feedstocks from a series of integrating catalytic
processing steps further support these facts. In the dawn of the
21st century (the eciency and economic competitiveness era),
we must redouble our eorts in expanding current technologies and approaches learning from our experience with simple
starting materials and aim to apply them to complex renewable
and waste feedstocks to pave the way to the widespread
implementation of bioreneries, where amazing happens!

Acknowledgements
RL gratefully acknowledges Ministerio de Ciencia e Innovacion,
Gobierno de Espana for the concession of a Ramon y Cajal
contract (RYC-2009-04199). Financial support through projects
MAT2010-21147 (MICINN) and P10-FQM-6711 (Consejeria
de Ciencia e Innovacion, Junta de Andalucia) is also gratefully
acknowledged.

References
1 A. J. Ragauskas, C. K. Williams, B. H. Davison, G. Britovsek,
J. Cairney, C. A. Eckert, W. J. Frederick, J. P. Hallett, D. J. Leak,
C. L. Liotta, J. R. Mielenz, R. Murphy, R. Templer and
T. Tschaplinski, Science, 2006, 311, 484489.
2 T. Werpy and G. Petersen, in Top value added chemicals from
biomass. Vol. 1. Results of screening for potential candidates from
sugars and synthesis gas, U.S. Dep. Energy, O. Sci. Tech. Inf.,
2004. http://www.nrel.gov/docs/fy04osti/35523.pdf.
3 J. J. Bozell and G. R. Petersen, Green Chem., 2010, 12, 539554.
4 J. Chheda, G. W. Huber and J. A. Dumesic, Angew. Chem., Int. Ed.,
2007, 46, 71647183.
5 U.S. National Science Foundation 2008, in Breaking the chemical
and engineering barriers to lignocellulosic biofuels: next generation
hydrocarbon bioreneries. http://www.ecs.umass.edu/biofuels/Images/
Roadmap2-08.pdf.
6 T. Okuhura, Chem. Rev., 2002, 102, 36413666.
7 R. Rinaldi and F. Schuth, Energy Environ. Sci., 2009, 2, 610626.
8 M. Toda, A. Takagaki, M. Okumura, J. N. Kondo, S. Hayashi,
K. Domen and M. Hara, Nature, 2005, 438, 178.
9 R. J. White, V. Budarin, R. Luque, J. H. Clark and
D. J. Macquarrie, Chem. Soc. Rev., 2009, 38, 34013418.
10 B. J. Arena, Appl. Catal., A, 1992, 87, 219229.
11 K. Lehnert and P. Claus, Catal. Commun., 2008, 9, 25432546.
12 J. C. Serrano-Ruiz, D. Wang and J. A. Dumesic, Green Chem.,
2010, 12, 574577.
13 Ethanol producer magazine, International ethanol report 2010,
http://www.ethanolproducer.com/articles/6696/international-ethanolreport-2010.

This journal is

The Royal Society of Chemistry 2011

Published on 29 June 2011. Downloaded by Universidad de Malaga on 28/10/2014 15:46:44.

View Article Online

14 M. Nia, D. Y. C. Leung and M. K. H. Leung, Int. J. Hydrogen


Energy, 2007, 32, 32383247.
15 J. Bedia, R. Barrionuevo, J. Rodr guez-Mirasol and T. Cordero,
Appl. Catal., B, 2001, 103, 302310.
16 Ethanol producer magazine, Braskem starts up ethanol-toethylene plant, http://www.ethanolproducer.com/articles/7022/
braskem-starts-up-ethanol-to-ethylene-plant.
17 A. Talalay and M. Magat, Synthetic Rubber from Alcohol,
Interscience, New York, 1945.
18 Method of producing propylene containing biomass-origin
carbon, European Patent Application EP1953129 (A1) to Mitsui
Chemicals Inc. (2008).
19 C. H. Christensen, J. Rass-Hansen, C. Marsden, E. Taarning and
K. Egeblad, ChemSusChem, 2008, 1, 283289.
20 Y. Tu and Y. Chen, Ind. Eng. Chem. Res., 1998, 37, 26182622.
21 C. H. Christensen, B. Jorgensen, J. Rass-Hansen, K. Egeblad,
R. Madsen, S. K. Klitgaard, S. M. Hansen, M. R. Hansen,
H. C. Andersen and A. Riisager, Angew. Chem., Int. Ed., 2006,
118, 46484651.
22 B. Jrgensen, S. E. Christiansen, M. L. D. Thomsen and
C. H. Christensen, J. Catal., 2007, 251, 332337.
23 (a) X. Tong, Y. Ma and Y. Li, Appl. Catal., A, 2010, 385, 113;
(b) A. A. Rosatella, S. P. Simeonov, R. F. M. Frade and
C. A. M. Afonso, Green Chem., 2011, 13, 754793.
24 Y. Roman-Leshkov, J. Chheda and J. A. Dumesic, Science, 2006,
312, 19331935.
25 H. Zhao, J. E. Holladay, H. Brown and Z. C. Zhang, Science,
2007, 316, 15971600.
26 Y. Roman-Leshkov, C. J. Barrett, Z. Y. Liu and J. A. Dumesic,
Nature, 2007, 447, 982986.
27 T. Thananatthanachon and T. B. Rauchfuss, Angew. Chem., Int. Ed.,
2010, 49, 66166618.
28 R. M. West, Z. Y. Liu, M. Peter and J. A. Dumesic, ChemSusChem, 2008, 1, 417424.
29 P. Maki-Arvela, B. Holmbom, T. Salmi and D. Yu. Murzin, Catal.
Rev., 2007, 49, 197340.
30 O. Casanova, S. Iborra and A. Corma, ChemSusChem, 2009, 2,
11381144.
31 M. L. Ribeiro and U. Schuchardt, Catal. Commun., 2003, 4, 8386.
32 Y. Y. Gorbanev, S. K. Klitgaard, J. M. Woodley, C. H. Christensen
and A. Riisager, ChemSusChem, 2009, 2, 672675.
33 N. K. Gupta, S. Nishimura, A. Takagaki and K. Ebitani, Green
Chem., 2011, 13, 824827.
34 R. Weingarten, J. Cho, W. C. Conner Jr. and G. W. Huber, Green
Chem., 2010, 12, 14231429.
35 S. Lima, P. Neves, M. M. Antunes, M. Pillinger, N. Ignatyev and
A. A. Valente, Appl. Catal., A, 2009, 363, 9399.
36 A. S. Mamman, J. M. Lee, Y. C. Kim, I. T. Hwang, N. J. Park,
Y. K. Hwang, J. S. Chang and J. S. Hwang, Biofuels, Bioprod.
Bioren., 2009, 2, 438454.
37 K. J. Zeitsch, The Chemistry and Technology of Furfural and Its
Many By-Products, 1st edn, Elsevier, Amsterdam, 2000, Sugar
Series, vol. 13, pp. 150155.
38 B. V. Timokhin, V. A. Baransky and G. D. Eliseeva, Russ. Chem.
Rev., 1999, 68, 7384.
39 S. Sitthisa and D. E. Resasco, Catal. Lett., 2011, 141, 784791.
40 S. Shi, H. Guo and G. Yin, Catal. Commun., 2011, 12, 731733.
41 R. Xing, A. V. Subrahmanyan, H. Olcay, W. Qi, G. van Walsum,
H. Pendse and G. W. Huber, Green Chem., 2010, 12, 19331946.
42 Y. Fan, C. Zhou and X. Zhu, Catal. Rev. Sci. Eng., 2009, 51, 293324.

This journal is

The Royal Society of Chemistry 2011

43 M. S. Holm, S. Saravanamurugan and E. Taarning, Science, 2010,


328, 602605.
44 J. C. Serrano-Ruiz and J. A. Dumesic, Green Chem., 2009, 11,
11011104.
45 A. Cukalovic and C. V. Stevens, Biofuels, Bioprod. Bioren., 2008,
2, 505529.
46 C. Delhomme, D. Weuster-Botz and F. E. Kuhn, Green Chem.,
2009, 11, 1326.
47 (a) T. J. Farmer, J. H. Clark and F. E. I. Deswarte, Biofuels,
Bioprod. Bioren., 2009, 3, 7290; (b) Y. Tokiwa and B. P. Caiabia,
Can. J. Chem., 2008, 86, 548555; (c) V. F. Wendisch, M. Bott and
B. J. Eikmanns, Curr. Opin. Microbiol., 2006, 9, 268274.
48 (a) M. P. Dorado, S. K. C. Lin, A. Koutinas, C. Y. Du,
R. H. Wang and C. Webb, J. Biotechnol., 2009, 143, 5159;
(b) R. Luque, S. K. C. Lin, C. Y. Du, A. Koutinas, R. H. Wang,
C. Webb and J. H. Clark, Green Chem., 2009, 11, 193200.
49 S. K. C. Lin, R. Luque, J. H. Clark, C. Webb and C. Du, Energy
Environ. Sci., 2011, 4, 14711479.
50 (a) V. L. Budarin, J. H. Clark, R. Luque, D. J. Macquarrie,
A. Koutinas and C. Webb, Green Chem., 2007, 9, 992995;
(b) V. Budarin, R. Luque, D. J. Macquarrie and J. H. Clark,
Chem.Eur. J., 2007, 13, 69146919.
51 (a) U. G. Hong, S. Hwang, J. G. Seo, J. Lee and I. K. Song, J. Ind.
Eng. Chem., 2011, 17, 316320; (b) R. Luque, J. H. Clark,
K. Yoshida and P. L. Gai, Chem. Commun., 2009, 53035305;
(c) R. M. Deshpande, V. V. Buwa, C. V. Rode, R. V. Chaudhari
and P. L. Mills, Catal. Commun., 2002, 3, 269274.
52 S. W. Fritzpatrick, World Patent 9640609 to Bione Incorporated,
1997.
53 I. T. Horvath, H. Mehdi, V. Fabos, L. Boda and L. T. Mika, Green
Chem., 2008, 10, 238242.
54 L. Deng, J. Li, D. M. Lai, Y. Fu and Q. X. Guo, Angew. Chem.,
Int. Ed., 2009, 48, 65296532.
55 J. J. Bozell, L. Moens, D. C. Elliott, Y. Wang, G. G. Neuenscwander,
S. W. Fritzpatrick, R. J. Bilski and J. L. Jarnefeld, Resour., Conserv.
Recycl., 2000, 28, 227239.
56 J. P. Lange, R. Price, P. M. Ayoub, J. Louis, L. Petrus, L. Clarke
and H. Gosselink, Angew. Chem., Int. Ed., 2010, 49, 44794483.
57 J. Q. Bond, D. Martin-Alonso, D. Wang, R. M. West and
J. A. Dumesic, Science, 2010, 327, 11101114.
58 (a) M. Pagliaro and M. Rossi, Future of glycerol, new usages for a
versatile raw material, RSC publishing, Cambridge, UK, 2008;
(b) http://www.biodieselmagazine.com/article.jsp?article_id=4080.
59 N. Luo, X. Fu, F. Cao, T. Xiao and P. P. Edwards, Fuel, 2008, 87,
34833489.
60 N. Ueda, Y. Nakagawa and K. Tomishige, Chem. Lett., 2010,
506507.
61 (a) A. Behr, J. Eilting, K. Irawadi, J. Leschinski and F. Lindner,
Green Chem., 2008, 10, 1330; (b) C. H. C. Zhou, J. N. Beltramini,
Y. X. Fan and G. Q. M. Lu, Chem. Soc. Rev., 2007, 37, 527549.
62 E. Taarning, S. Saravanamurugan, M. S. Holm, J. Xiong,
R. W. West and C. H. Christensen, ChemSusChem, 2009, 2,
625627.
63 R. D. Cortright, R. R. Davda and J. A. Dumesic, Nature, 2002,
418, 964967.
64 (a) N. Ji, T. Zhang, M. Zheng, A. Wang, H. Wang, X. Wang and
J. G. Chen, Angew. Chem., Int. Ed., 2008, 47, 85108513; (b) J. Sun
and H. Liu, Green Chem., 2011, 13, 135142.
65 E. L. Kunkes, D. A. Simonetti, R. M. West, J. C. Serrano-Ruiz,
C. A. Gartner and J. A. Dumesic, Science, 2008, 322, 417421.

Chem. Soc. Rev., 2011, 40, 52665281

5281

S-ar putea să vă placă și