Download as pdf or txt
Download as pdf or txt
You are on page 1of 20

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

CRITICAL REVIEW www.rsc.org/greenchem | Green Chemistry

Glycerol dehydration to acrolein in the context of new uses of glycerol


Benjamin Katryniok,a,b Sébastien Paul,a,b,c Virginie Bellière-Baca,d Patrick Reye and Franck Dumeignil*a,b
Received 9th July 2010, Accepted 5th October 2010
DOI: 10.1039/c0gc00307g
Published on 12 November 2010. Downloaded by Pennsylvania State University on 17/09/2016 04:41:30.

Catalytic dehydration of glycerol to acrolein has the potential to valorise the glut of crude glycerol
issuing from biodiesel production. This reaction requires catalysts with appropriate acidity, and
intensive research activities have been focused on the application of families of catalysts including
zeolites, heteropolyacids, mixed metal oxides and (oxo)-pyrophosphates, as their acidic properties
are well-known. Nevertheless, their deactivation by coking remains the main obstacle in the way of
large-scale industrial applications. Considering this important issue, various technologies have been
proposed for regenerating the catalysts. This review shows that a well-balanced combination of an
appropriate catalytic system together with an adapted regeneration process could put large-scale
industrial applications within reach.

1 Introduction The capacity of biodiesel production is expanding all over


the world. For example, in 2008, the USA produced 2.3 million
The finite reserves of fossil-fuel feedstocks have encouraged tonnes of biodiesel, while the EU produced 7.8 million tonnes
intensive research activities for developing substitutes/additives (Fig. 1), and these values are expected these to double by 2012.4,5
for fuels such as biodiesel, bioethanol or biokerosene. Due to This growth is accompanied by a significant increase in glycerol
their origin from biomass, they have the strong advantage of production, as this a significant by-product (~10 wt%) of the
a lower carbon footprint than fuels derived from fossil-fuel biodiesel process (Scheme 1). Simple projections enable one to
resources. However, from an economic point of view, it is not forecast that 1.54 million tonnes of glycerol will be generated
yet possible to produce any of these biofuels competitively. worldwide in 2015,6 all of which will have to be efficiently
Nevertheless, political decisions have pushed the production of processed in order to achieve a sustainable business.
fuels from bioresources in order to be able to fulfill the CO2
reduction objectives fixed by the Kyoto climate protocol.1 For
example, the European Union has planned a progressive increase
in the mandatory proportion of bioethanol blended in gasoline,
and of biodiesel blended in diesel fuel. The blended amounts
must reach 10% and 7% by 2015, respectively.2
The alternative gasoline production route using vegetable
oils and fats has attracted the attention of many academic
and industrial researchers.3 The raw materials for biodiesel
production are vegetable oils and fats – from canola, soy, corn,
etc. – and a mono-alcohol (usually methanol), which is used to
cleave the fatty acids from their glycerol backbone to yield the Fig. 1 Comparative evolution of the quantities of biodiesel produced
desired fatty acid esters (Scheme 1). These esters can be directly in the EU and the US.5,7
used as biodiesels, but they are commonly blended with diesels
derived from fossil-fuels to meet the regulations. Depending on the process used for the cleavage of the fatty
acids, the purity of the crude glycerol can vary greatly. The
crude glycerol obtained from most of the conventional biodiesel
processes contains ca. 80 wt% of glycerol, but it also contains
water, methanol, traces of fatty acids as well as various inorganic
and organic compounds (called ‘MONG’, denoting ‘Matter
Scheme 1 Transesterification reaction of vegetable oils to yield Organic Non-Glycerol’) (Table 1).
biodiesel. As a consequence, in most of the cases, crude glycerol must
be purified by an expensive distillation step prior to further
use, to meet the requirements of the downstream processes.
a
Univ. Lille Nord de France, F-59000, Lille, France. The proportion of glycerol that is refined is actually steadily
E-mail: Franck.Dumeignil@univ-lille1.fr; Fax: +33 (0)3.20.43.65.61; decreasing, due to the high cost of the distillation step, together
Tel: +33 (0)3.20.43.45.38 with a rapid growth of the quantity of crude glycerol produced,
b
CNRS UMR8181, Unité de Catalyse et Chimie du Solide, UCCS,
F-59655, Villeneuve d’Ascq, France and also (primarily) because of the absence of any market
c
ECLille, F-59655, Villeneuve d’Ascq, France able to absorb the massive overproduction (Fig. 2). In fact, an
d
RHODIA France, 52 Rue de la Haie Coq, 93308, Aubervilliers, France increase in the price of glycerol – together with a sustainable and
e
ADISSEO France SAS, Antony Parc 2-10, 92160, Antony, France significant demand – would automatically result in a decrease in

This journal is © The Royal Society of Chemistry 2010 Green Chem., 2010, 12, 2079–2098 | 2079
View Article Online

Table 1 Examples of qualities of glycerol derived from some biodiesel


production processes.8 Compositions are given in wt%

Company production site

Robbe/Diester Saipol/Diester Diester/Bioenergy


Compiègne (France) Rouen (France) Marl (Germany)

Glycerol 65 93 85
Published on 12 November 2010. Downloaded by Pennsylvania State University on 17/09/2016 04:41:30.

Water 31 4 10
MONGa <1 <1 <0.5
Salts 3 (Na2 SO4 ) 2.5 (NaCl) 4.5 (NaCl)
Methanol 0.3 0.2 <0.01
Fig. 3 Distribution of the glycerol consumption by sector/
a
MONG = Matter Organic Non-Glycerol.
application.10

such when the amount of water has to be controlled, namely in


glue or other adhesives. Further, the presence of hydroxyl groups
leads to the formation of intra- and inter-molecular hydrogen
networks, which explains its high boiling point (290 ◦ C at
atmospheric pressure) and high viscosity. This latter rheological
property is at the origin of the use of glycerol as a softener
in resins and plastics but also as a lubricant, for instance in
pharmaceutical applications. In addition, glycerol is non-toxic
and has a sweet taste. It can thus be incorporated in food,
medicines and cosmetics. However, as mentioned earlier, the
Fig. 2 Global crude glycerol production compared to the amount crude glycerol from the biodiesel processes contains impurities,
distilled. ‘Delta’ is the amount of glycerol that is not upgraded, and and is therefore not suitable for such applications without a
usually burned.6,7 preliminary purification stage. In addition, the size of the existing
market is not sufficient to absorb the huge amount of glycerol
the production costs of the transesterification process, which currently produced, and the gap between the absorption capacity
up to now had to be entirely compensated for by the sales of the market and the amount of glycerol produced will increase
of biodiesel.9 New economical ways of using glycerol must in the near future if no new applications are found.
therefore be developed in order to substantially increase the Today, the crude glycerol that is not refined is generally
demand and the price of crude glycerol, and therefore to ensure burned (Fig. 2), which must be considered as a tragic waste
the sustainability of the biodiesel sector.10 of a potentially very useful organic raw material. Glycerol is
More than 1500 direct applications of glycerol are already a molecule with a large functionalisation potential that offers
known, especially in cosmetics, pharmaceuticals and food numerous opportunities for chemical or biochemical conver-
industries.11 The large versatility of glycerol use is based on sions for producing value-added chemicals. A non-exhaustive
both its chemical and physical properties (Fig. 3). Due to the selection of these possibilities is shown in Scheme 2, and is
presence of three hydroxyl groups, glycerol is completely soluble further discussed below.
in water and alcohols, whereas it is completely insoluble in For the moment, only a few applications utilise glycerol on a
hydrocarbons. It is a very hydrophilic species, and is employed as large scale. One of them is the halogenation to epichlorhydrin,

Scheme 2 Selected glycerol valorisation target molecules.

2080 | Green Chem., 2010, 12, 2079–2098 This journal is © The Royal Society of Chemistry 2010
View Article Online

which is an important intermediate for epoxy resins. The process sidering the petrochemical feedstock depletion issues, sustain-
uses hydrochloric acid in the presence of organic acids (caprylic able resources will become more and more competitive, not to
acid – Solvay/acetic acid – Dow)12 as catalysts working in the mention their positive effect in terms of impact on the climate.
gaseous phase at 180–220 ◦ C under a pressure of 1–5 bar.13 Within this context, a massive bioresourced and sustainable
This technology was commercialised in 2007 by Solvay, which acrolein production is an important challenge, which has been
operates an existing production facility in France that was accepted by academic and industrial researchers. An economics
formerly used to produce glycerol from epichlorhydrin. Another study has shown that competitive production of acrolein from
Published on 12 November 2010. Downloaded by Pennsylvania State University on 17/09/2016 04:41:30.

commercialised process for large-scale consumption of glycerol glycerol could be possible if the price of glycerol became less
is the reforming over Pt–Rh catalysts to yield syngas. This than ca. 300 US$/t.25 In January 2010, refined glycerol still cost
latter can either be used in the Fischer–Tropsch process for 500–550 US$/t, but crude glycerol was only around 100 US$/t.
the synthesis of alkanes (Biomass to Liquid, BtL) or for the This makes this latter a potentially very competitive raw material
synthesis of methanol as performed by BioMCN in Delfzijl for acrolein production, even though some technical difficulties
(Netherlands) with a capacity of 200 kt/year.14–16 From this in the application of crude glycerol have still to be solved, in
technology is derived the sustainable production of hydrogen particular avoiding catalyst poisoning by the impurities.26
– an energy vector with an expected high potential in the near Due to its toxicity, acrolein is usually directly converted to
future – through the water-gas shift process. As another option, the desired high value-added derivatives. Most acrolein is used
the etherification to glycerol tert-butyl ether (GTBE) targets for the synthesis of acrylic acid, which is used, for example, as a
the classical petrol-based economy. Indeed, this compound starting material for synthesising sodium polyacrylate (Scheme
is an environmentally friendly additive in gasoline, and can 4). According to its physical properties, this polymer, classified as
beneficially substitute for the problematic methyl tert-butyl ether a superabsorbent, finds uses in hygiene products, such as diapers.
(MTBE). Nevertheless, the associated process is still far from The annual worldwide market for this superabsorbent polymer
reaching industrial applications.17 (SAP) is estimated at 1.9 million tonnes for 2010.27
The selective reduction and esterification of glycerol are
two processes with potential impacts on glycerol consumption
at a medium scale. The former can be used to yield either
propylene glycol (MPG) or 1,3-propanediol (PD), which are
valuable intermediates in the polymer industry. The latter targets
the esters of glycerol – namely, monoacylglycerol (MAG) and
diacylglycerol (DAG) – that find applications as emulsifiers, Scheme 4 Industrial synthesis of sodium polyacrylate, a superab-
e.g., in food (margarines and sauces) or in cosmetics. This sorbent polymer.
process can either be catalysed by a conventional catalyst or by
enzymes (lipase-type).18 Recently, the production of MPG has The second largest consumer of acrolein is represented by the
been commercialised by Synergy Chemicals with a production synthesis of DL-methionine via 3-methylthio-propionaldehyde
capacity of 30 kt/year.10,19 as an intermediate (Scheme 5).28 DL-Methionine is an essential
A small-scale application, but aiming at the synthesis of amino-acid, which cannot be synthesised by living organisms,
high-value fine chemicals, consists on the partial oxidation of and is widely used in meat production to accelerate animal
glycerol to carboxylic acids (i.e., glyceric acid, tartronic acid and growth. The annual worldwide production capacity of DL-
ketomalonic acid), aldehydes (i.e., glyceraldehyde) or ketones methionine is about 500 000 tonnes.29 As natural methionine
(i.e., dihydroxyacetone). The main challenge is to find oxidation sources like plants and microorganisms provide concentrations
catalysts that are highly selective to the target molecule, among that are too low, it has to be synthesised on the industrial scale.
a large quantity of possible reaction products. One example is It is estimated that the global demand will increase by 3–7% in
the selective oxidation to dihydroxyacetone over Bi/Pt catalysts, the near future.30
giving 37% yield at 70% conversion of glycerol.20 Oxidation
can also be performed using genetically modified bacteria or
electrochemical reactions.21,22
However, one of the most promising routes to glycerol
valorisation lies in its catalytic dehydration to acrolein, which
is an important industrial intermediate for the chemical and the
agro-industries. The synthesis of acrolein is currently based on
the selective oxidation of propene over complex multicomponent
BiMoOx -based catalysts (Scheme 3). The selectivity reached in
this process is close to 85% at 95% conversion.23 New approaches
with propane as a starting material are currently being explored Scheme 5 Chemical pathways for the industrial manufacture of DL-
at the laboratory scale, but they suffer from insufficient yields, methionine.
which are incompatible with further commercialisation.24 Con-
In the following sections, we will present the state-of-the-
art on the sustainable production of acrolein by gas-phase
dehydration of glycerol. In fact, since our last review on the
Scheme 3 Catalytically-assisted selective oxidation of propene. dehydration of glycerol,31 the number of publications in this

This journal is © The Royal Society of Chemistry 2010 Green Chem., 2010, 12, 2079–2098 | 2081
View Article Online

field has greatly increased (Fig. 4). In particular, the funda- They reproduced the experiments over the lithium phosphate
mental scientific aspects of the reaction have been illuminated. catalyst and compared its performance to those of a set of acid
Therefore, we will first discuss the general properties needed to catalysts with well defined Hammett acidities (HAs) ranging
give high catalytic performances, and then we will describe in from +3 to -8.2. They claimed that aqueous solutions from 10
detail the three most common types of catalytic systems used to 40 wt% of glycerol could be converted at 300 ◦ C over alumina-
in this reaction. The possibility of catalyst regeneration, such supported phosphorous acid, to yield acrolein with a selectivity
as the dehydration of glycerol in the presence of oxygen, will up to 75% at full glycerol conversion. The main by-product was
Published on 12 November 2010. Downloaded by Pennsylvania State University on 17/09/2016 04:41:30.

also be commented upon. After a short look at liquid-phase acetol (hydroxyacetone), with a selectivity close to 10%. Neher
dehydration, the fundamentals of the reaction will be discussed et al. also pursued research on alumina-supported phosphorous
in a section devoted to the insights in the understanding of the acid, but with a view to producing 1,2- and 1,3-propanediol.37
mechanism. Finally, the possibility of using crude glycerol will The yield of acrolein formed as an intermediate was, however, a
be discussed, before concluding the discussion and giving an little lower (70.5%).
outlook on the possible directions for further research efforts.
3 Acidity of the catalysts
Clearly, the acidity of the active phase is a crucial parameter
that influences the catalytic performance and stability. The first
quantitative studies on this effect were published by Dubois
et al. in 2006.38,39 They carried out several tests with zeolites,
NafionR , heteropolyacids (HPAs) and also with different types
of acid-impregnated metal oxides. All these catalysts have a well
defined HA ranging from -9 to -18. The selectivity for acrolein
depended on the catalyst: For NafionR catalysts (HA = -12)
Fig. 4 Evolution of the number of publications dealing with glycerol and for tungsten oxide on zirconium oxide (HA = -14.5), the
dehydration to acrolein (plotted using data collected using ScopusTM by selectivity for acrolein was close to 70% at full conversion, while
performing a search with “glycerol, dehydration, acrolein” as keywords). the selectivity over zeolitic catalysts (HA < +2) did not exceed
60%. The authors concluded that catalysts with an HA of -10
to -16 were the best candidates for the selective dehydration of
glycerol to acrolein.
2 Historical context
Inspired by these first results, a second research group focused
It has been well known for a long time that heating glycerol its activities on the influence of catalyst acidity. In 2007, Chai
induces its decomposition to acrolein and water, with a certain et al. published two papers on the sustainable production
amount of by-products. However, an acid catalyst is needed to of acrolein from glycerol. Whereas the first paper focused
provide good control of the reaction and to obtain a significant on niobium oxide as a catalyst,40 the second compared the
yield of the unsaturated aldehyde at a moderate temperature. catalytic performances of different types of acid catalysts.41
The first patent on this subject was published by Schering in As the acidity of niobium oxide is inversely related to the
1930 in France.32 The reaction was carried out in the gaseous calcination temperature of the hydrous niobium oxide precursor
phase over supported lithium phosphate as a catalyst, with the (as previously described by Chen et al.42 and Lizuka et al.43 ),
yield of acrolein being close to 75%. Later, in 1934, Groll and this compound is particularly suitable as a model catalyst for
Hearne from the Shell company applied for a patent for the studying the influence of acidity on a given reaction. The best
dehydration of an aqueous glycerol solution in the presence of results were obtained for catalysts calcined at low temperature
sulfuric acid at 190 ◦ C (supposedly the boiling point of the (250–300 ◦ C). In addition to their high acidity, these catalysts are
solution).33 The product was recovered in the gas phase with a also advantageous as they have higher specific surface area than
yield of nearly 50%. Dehydration in the liquid phase was further the catalysts calcined at higher temperatures. Nevertheless, the
studied and patented by Hoyt and Manninen, this time using selectivity for acrolein barely exceeded 50% at 90% of glycerol
a heterogeneous catalyst based on phosphorous acid supported conversion. In addition, the catalysts calcined at low temperature
on clay.34 It is interesting to note that these authors even at that showed high carbon deposition, and were therefore subject to
time preferred petroleum oil as a reaction medium, due to its quick deactivation. From these observations, it was already
high boiling point of more than 300 ◦ C. Thereby, they could use possible to conclude that the strength of the acid catalyst must
higher reaction temperatures than Groll and Hearne,33 who were be carefully kept in a narrow range, because acids that are too
limited to 190 ◦ C, as the reaction was performed in the aqueous weak give low selectivity, whereas acids that are too strong result
phase. This combination of a heterogeneous catalyst and a high in accelerated deactivation.
boiling point reaction medium allowed them to obtain a yield of Therefore, Chai’s team screened numerous catalysts, which
72.3%. However, these early works – in both the gaseous and the were classified into four different groups according to their
liquid phases – remained undeveloped until the end of the 20th acidity.41 Group 1 contained basic catalysts with an HA higher
century, when cheaper glycerol from the biodiesel production than +7, like magnesium oxide. These catalysts showed no
process became available. selectivity for acrolein at all. Group 2 contained catalysts with
In 1994 and 1995, Neher et al. published two patents35,36 as an HA between -3 and +7, like zirconium oxide. According to
a follow-up of the pioneering work of Schering-Kahlbaum.32 the work of Neher,35 these catalysts should give high acrolein

2082 | Green Chem., 2010, 12, 2079–2098 This journal is © The Royal Society of Chemistry 2010
View Article Online

selectivities, but, in practice, the selectivity did not exceed 30%, According to Alhanash et al.44 the reaction over Brønsted-
while the performances remained quite stable for 10 h on stream. type acid catalysts starts with the protonation of the secondary
Group 3, which was more promising, contained catalysts with an hydroxyl group of glycerol by a proton from a Brønsted site, as
HA between -8 and -3. In this group, we find alumina-supported postulated by Bühler for homogenous reaction media (Scheme
phosphorous acid next to alumina-supported HPAs, niobium 6a).87 The resulting intermediate is transformed by release of a
oxide (calcined at 400–500 ◦ C), as well as HZSM zeolite and hydronium ion (H3 O+ ) followed by a keto–enol rearrangement to
pure alumina. In good agreement with previous results,38,39 the give 3-hydroxypropionaldehyde and – after a second dehydration
Published on 12 November 2010. Downloaded by Pennsylvania State University on 17/09/2016 04:41:30.

selectivity was generally higher than that observed for Group 2 step – acrolein. The reacted Brønsted acid site can be regenerated
catalysts, with the exception of pure alumina and niobium oxide (re-protonated) by the aforementioned released hydronium ion.
calcined at 500 ◦ C. Interesting results were also observed over The reaction over Lewis acid sites proceeds in a completely
the alumina-supported phosphotungstic HPA and over a mixed different way (Scheme 6b), as the glycerol interacts via concerted
phase of tungsten oxide/zirconium oxide, with approximately transfer of a terminal hydroxyl group on one of the metal centers
70% selectivity at 70% conversion in both cases after 2 h on- and migration of the secondary proton to the other metal center.
stream. Unfortunately, these catalysts showed poor stability, Consequently, 2,3-dihydroxypropene and pseudo-Brønsted sites
and their performance significantly decreased during time on- are formed. The enol will tautomerise to acetol, whereas
stream (for both catalysts, the glycerol conversion dropped from the pseudo-Brønsted site can either catalyse the dehydration
68–69% after 1 h under reactant flow to 23–25% at 10 h, the reaction, as in the case of the aforementioned Brønsted sites,
selectivity being more-or-less unaffected at roughly 66–70%). or can regenerate the initial Lewis site by thermal dehydration.
The other catalysts were on the whole less efficient, and gave Thereby, one can explain why Lewis acid sites generally show
selectivities of 35–55% at conversions of 55–100% after 10 h on- higher selectivity for acetol than Brønsted acid sites.
stream. In Group 4 were included catalysts with a HA less than
-8, like Hb-zeolite, niobium oxide calcined at 350 ◦ C, and also
alumina-silicate as well as sulfonated zirconium oxide. These
catalysts were less selective to acrolein than those of Group 3,
but their performances were less altered with time on-stream.
As an example, the glycerol conversion over an aluminosilicate
dropped only from 94% after 1 h under reactant flow to 75%
after 10 h, the acrolein selectivity being more-or-less unchanged
at roughly 43–46%. Nevertheless, like niobium oxide, Group 4
catalysts were also subject to detrimental coke formation, with
a carbon deposit of 100–400 mg per gram of catalyst after ten
hours on-stream.
In addition to the acidic strength of the catalyst, the type of the
acid sites present at their surfaces has an important influence on
their catalytic performance. When Chai et al. classified the cata-
lysts according to their strength,41 they mixed up Brønsted and
Lewis acid catalysts without differentiation. Whereas Brønsted
acids donate a proton (as for the aforementioned silicotungstic
acid, protonated zeolites or phosphorous acid), Lewis acids
Scheme 6 Reaction mechanism over basic catalysts, as proposed by
are acceptors for electron pairs (as for niobium oxide or pure
Kinage et al.44
alumina). From the catalytic results obtained by Chai et al.
over these two different types of acids, one can conclude that
they do not obey the same reaction pathway, as the authors As catalysts usually contain both acidic and basic sites,
systematically obtained rather low selectivities for acrolein over the influence of the basicity must also be considered. Kinage
Lewis acids.41 A more detailed study on the different catalytic et al. investigated the selective dehydration of glycerol to acetol
behaviors of Brønsted and Lewis acids in this reaction was and proposed a reaction mechanism over basic catalysts.46 In
proposed by Alhanash et al.,44 who compared a pure Brønsted contrast to the mechanisms proposed over acid catalysts, the
acid catalyst (acidic caesium salts of phosphotungstic acid) to reaction over basic catalysts does not begin with a dehydration
a pure Lewis acid catalyst (tin-chromium mixed oxide). They step but with a dehydrogenation step (Scheme 7). The resulting
showed that: 2,3-dihydroxypropanal can either further react via consecutive
∑ Lewis acid catalysts need a higher reaction temperature to dehydration and hydrogenation to yield acetol, or can form – via
be activated, due to a higher activation energy compared to a retro-aldol reaction – formaldehyde and hydroxyacetaldehyde,
Brønsted acid catalysts; which can be subsequently hydrogenated to ethylene glycol. With
∑ Lewis acid catalysts give a larger selectivity for ace- this reaction mechanism, one can explain the high selectivity for
tol, which is the major by-product observed during glycerol acetol observed by Chai et al. over basic catalysts.41
dehydration. In this section, we have seen that the strength and the type of
These results were confirmed by Kim et al., who compared the catalyst acidity are of utmost importance for obtaining high
HZSM-5 aluminosilicates with different ratios of alumina and catalytic performances. In the following sections, we will further
silica.45 discuss in detail the three most widely used catalytic systems for

This journal is © The Royal Society of Chemistry 2010 Green Chem., 2010, 12, 2079–2098 | 2083
Published on 12 November 2010. Downloaded by Pennsylvania State University on 17/09/2016 04:41:30. View Article Online

Scheme 7 Reaction mechanism over protonated zeolite (H-MFI) catalyst as proposed by Yoda et al.46

the dehydration of glycerol: supported HPAs, zeolites, and the grafted with zirconia before impregnation of the active phase.50
family of metal oxides, phosphates and pyrophosphates. We showed that the strong electronic interaction between the
zirconia and the heteropolyanion results in a decreased strength
of the Brønsted acid sites, whereby the long-term catalytic
4 Supported heteropolyacids performances were increased in comparison to bare silica (69%
The use of supported inorganic acids such as phosphorous acrolein vs. 24% respectively, after 24 h on-stream) due to less
acid or HPAs is one possibility to reach high performances coking. Nevertheless, it is worth mentioning that the quantity of
in the dehydration of glycerol to acrolein, as shown by Neher grafted zirconia and active phase necessitates careful balancing
and Chai.35,36,41 However, when using supported catalysts, new in order to avoid the introduction of undesired Lewis-acid
parameters in addition to acidity must be taken into account, character into the catalyst, which results in decreased selectivity
namely the influence of the support on the active-phase prop- for acrolein as shown by Alhanash et al.44
erties, the dispersion extent of the active phase on the support, The influence of the pore size distribution of the support was
and the distribution of the pore size of the support. discussed for the first time in 2007 by Tsukuda and coworkers,51
In 2008 and 2009, Chai et al. published two papers dealing who worked with silica-supported acids. They focused their
with zirconia- and silica-supported HPA catalysts.47,48 These study on the influence of the textural properties of the porous
studies showed that the nature of the catalyst support has a support on the selectivity, and used three commercial silicas
significant effect on the thermal stability and on the dispersion with pore sizes of 3, 6 and 10 nm, namely CARiACT Q3, Q6
of the Keggin-type active phase. The use of zirconia as a support and Q10. The active phase was either a HPA or a conventional
led to better results, as far as the deactivation of the catalyst inorganic acid like phosphorous or boric acid. Good results
is concerned. The Keggin-anion density at the surface of the were obtained using Keggin-type HPAs such as phosphotungstic
support was identified as being key for tuning the activity and acid (H3 PW12 O40 ) or silicotungstic acid (H4 SiW12 O40 ) deposited
the selectivity of the HPA for acrolein production, which is on the 6 nm silica support. At full conversion, the selectivity
in good agreement with the conclusions of Ning et al., who for acrolein was 65% and 74%, respectively, whereas the
used activated carbon-supported silicotungstic acid catalysts.49 molybdenum-based homologous HPA (i.e., H3 PMo12 O40 ) did
The latter claimed that a 10 wt% silicotungstic acid-supported not yield high selectivity (34%) (Table 2). When a catalyst based
catalyst gives the best acrolein space-time yield ever reported on silicotungstic acid over silica with 10 nm pore diameter was
in the literature (namely 68.5 mmol acrolein h-1 g-1 ), but this used, the selectivity could be further increased to 86%, whereas
value is misleading, as the calculation is based on the mass of for a support with smaller pores (namely 3 nm) the selectivity
active phase and not on the mass of catalyst. This promising slightly decreased to 67%. Furthermore, it was observed that
performance is attributed to the good dispersion of the HPA the small pore size of 3 nm resulted in a decreased catalytic
on the support surface and also to the relative quantities of activity (55% conversion), which most probably originates from
strong acid sites. Our team also recently used silica-supported enhanced coking resulting from steric limitations within the
silicotungstic acid as a catalyst, but on supports that were catalyst framework. The influence of temperature was also

Table 2 Catalytic results obtained over various supported heteropolyacidsa

Catalyst T react (◦ C) mcat (g) TOS (h) C (%) Sacrolein (%) Sacetol (%) Y acrolein (%) STY (mmol g-1 h-1 ) Ref.

15 wt% PW/SiO2 315 0.3 10 18 57 7 10 0.6 47


15 wt% PW/ZrO2 315 0.63 10 76 76 12 58 1.6 47
10 wt% SiW/AC 330 0.38 5 93 67 8 62 11.1 49
20 wt% PW/Al2 O3 275 0.3 5 99 52 8 51 5.1 52
20 wt% SiW/Al2 O3 275 0.3 5 98 64 8 63 6.3 52
30 wt% PMo/Q6 325 0.3 5 98 65 7 64 3.8 51
30 wt% PW/Q6 325 0.3 5 100 33 5 33 2.0 51
30 wt% SiW/Q6 325 0.3 5 100 74 7 74 4.4 51
30 wt% SiW/Q10 275 0.3 5 98 86 7 84 5.1 51
CsPW 275 0.3 1 100 98 nc 98 49.0 44
a
T react = reaction temperature; mcat = mass of catalyst; TOS = time on stream; C = conversion of glycerol; Sacrolein = selectivity for acrolein; Sacetol =
selectivity for acetol; Y acrolein = yield of acrolein; STY = space-time yield; nc = not communicated; PW = H3 PW12 O40 ; SiW = H4 SiW12 O40 ; CsPW =
Cs2.5 H0.5 PW12 O40 ; AC = active carbon; Q6 = CARiACT-Q6; Q10 = CARiACT-Q10.

2084 | Green Chem., 2010, 12, 2079–2098 This journal is © The Royal Society of Chemistry 2010
View Article Online

studied, 275 ◦ C being optimal in terms of acrolein yield, which that they were reduced during the reaction. When oxygen was
is consistent with earlier studies on Brønsted acid catalysts. co-fed, the reduction degree of vanadium was limited to 4+
Similar results were obtained by Atia et al., who investigated and that of molybdenum to 5+. Furthermore, the coking was
the catalytic performance of the silicotungstic acid supported slowed down, and the authors claimed that the carbon was
on alumina with 5 nm and 12 nm pore sizes.52 The reaction preferentially deposited on the vanadium centers. The carbon
conditions for the catalytic tests were chosen such as to match species constituting the coke layer was also analysed. According
those used by Tsukuda et al.,51 for the sake of easier comparison. to the EPR signal, the carbon was mainly in the sp3 hybridisation
Published on 12 November 2010. Downloaded by Pennsylvania State University on 17/09/2016 04:41:30.

The selectivity for acrolein increased from 65% for the alumina state, which corresponds to non-double-bonded C–C chains, like
support with a 5 nm pore size to over 85% for the alumina from the radical polymerisation of the C C bond in acrolein.
support with a 12 nm pore size, both at full conversion and Furthermore, if oxygen was present in the feed, the formation
for catalysts loaded with 20 wt% of active phase (Table 2). of C O carbonyl groups was observed.
Furthermore, Atia et al. found the same optimum reaction Another idea for using heteropolycompounds was inves-
temperature as Tsukuda, since the alumina-based catalysts also tigated by Dubois et al., who used salts of HPAs ob-
showed complete conversion for a temperature higher than tained by complete or partial substitution of the protons
275 ◦ C. of HPAs by metal cations.57 They used silicotungstic or
These results show that the size of the pores has a direct phosphotungstic acid as bases, and prepared their respec-
influence on the selectivity of the catalysts. If steric limitations – tive salts using elements such as caesium, rubidium, calcium
like pores that are too narrow – hinder the rapid desorption or bismuth, or transition metals like zirconium, lanthanum,
and diffusion of the reactants and of the products in the iron or hafnium. The partially neutralised caesium-containing
porous network of the catalyst, then condensation is more likely HPAs (Cs2.5 H0.5 PW12 O40 and Cs2.5 H1.5 SiW12 O40 ) and rubidium-
to occur, which results in lower selectivity for acrolein and containing HPAs (Rb2.5 H0.5 PW12 O40 ) required significantly lower
deactivation of the catalyst due to carbon deposition (coke). reaction temperatures (260–280 ◦ C vs. 300–350 ◦ C) than the
From the aforementioned studies, one can identify two totally proton-free compounds, and further showed higher
important parameters that govern the catalytic performances selectivity for acrolein in oxydehydration (90% vs. 50–70%). The
of supported inorganic acids: (i) the dispersion of the active best result was reported for Cs2.5 H1.5 SiW12 O40 , with a 93.1% yield
phase and (ii) the textural properties of the support. These two of acrolein. Nevertheless, when this catalyst was used without
parameters were further investigated by Chai et al. and Sato co-feeding of oxygen the catalytic performance was much lower,
et al., who focused on supported HPAs and on their catalytic with an acrolein yield of no more than 40%. Comparable results
performances with regard to the textural properties of the porous with caesium salts of phosphotungstic acid were reported by
support.53–55 They carried out a huge number of performance Alhanash et al., who also obtained acrolein yields of up to
tests, by varying both the loading of active phases and the 98% during the first hour of reaction, which is the highest
nature of the supports. Here also, phosphomolybdic acid was performance ever reported, with a productivity up to 49 mmol
unselective compared to phosphotungstic acid and silicotungstic g-1 h-1 (Table 2).44 Nevertheless, these authors admit that the
acid.51 With 20–30 wt% silicotungstic acid deposited on two catalyst deactivates rapidly, and that excellent performances
porous silicas with pore sizes of 6 and 10 nm, respectively, they claimed were observed only for the first few hours of reaction.
obtained a selectivity for acrolein up to ca. 87% at 275 ◦ C in In conclusion, all these results suggest that the use of
both cases. Smaller pore diameters (3 nm) or a higher reaction Keggin-type heteropolycompounds is a highly promising means
temperature (300 ◦ C) always led to a decrease in the catalytic of obtaining high catalytic performances. They offer many
performance. It is worth mentioning that the influence of the possibilities for tuning their acidity due to the possible variations
BET specific surface area of the support was also studied, but in their composition, which can be adjusted by modifying the
this proved to have only a small effect. central atom, the addenda atoms and the counter-ions. When
Whereas Chai and Tsukuda51,53 showed that silicotungstic these compounds are deposited on a support, it is also possible
acid and phosphotungstic acid gave the best performances, to adjust an additional parameter, as one can even control the
due to their higher acidities, the group of Brückner et al.56 pore size of the catalyst, and thus reduce diffusion limitations.
recently decided to focus their investigations on supported phos-
phomolybdic (H3 PMo12 O40 ) and the vanado-phosphomolybdic
5 Zeolites
(H4 PMo11 VO40 ) acids. Using electron paramagnetic resonance
spectroscopy (EPR), they showed that the Keggin structure The use of supported heteropolycompounds is not the only way
of the molybdenum compounds was partially destroyed upon of obtaining good catalytic performances. Zeolitic structures are
impregnation on silica-alumina supports and that further promising, and have been well studied since the work of Dubois
calcination at 350 ◦ C led to total decomposition to simple et al.39
metal oxides. These catalysts were then tested in the glycerol In addition to ZSM-5 and b-zeolite (already patented by
dehydration reaction, and the oxidation state of the metal atoms Dubois et al.39 ), Li and Zhuang et al. studied other zeolite
was followed by Operando EPR. The catalytic performance was structures such as MCM-49, MCM-22, MCM-56 and ZSM-
close to that already observed for these compounds by Chai 11.58,59 The obtained acrolein yields were in the range of 70–85%
et al.,53 that is to say quite poor (the selectivity for acrolein did for all the catalysts, with nearly no decrease in the catalytic
not exceed 25%). On the other hand, the Operando analysis performance during 400 h of reaction.
of the catalyst revealed that the oxidation state of the metals In 2007, Okuno and coworkers patented metallosilicate
was 3+ for vanadium and 4+ for molybdenum, which indicates catalysts with an MFI structure, a typical structure of zeolites

This journal is © The Royal Society of Chemistry 2010 Green Chem., 2010, 12, 2079–2098 | 2085
View Article Online

Table 3 Catalytic performances observed over various zeolitesa

Catalyst T react (◦ C) mcat (g) TOS (h) C (%) Sacrolein (%) Sacetol (%) Y acrolein (%) STY (mmol g-1 h-1 ) Ref.

b-Zeolite 300 4.2 1.5 100 57 10 57 5.8 39


HZSM-5 300 6.4 1 79 49 6 39 2.6 39
MFI+Ga 360 9 nc 96 62 nc 59 4.6 60
MFI (28)b 360 9 nc 92 69 nc 63 4.8 61
Published on 12 November 2010. Downloaded by Pennsylvania State University on 17/09/2016 04:41:30.

MFI (200)b 360 9 nc 100 61 nc 61 4.6 61


ZSM-5 (30)b 315 0.3 2 52 42 7 22 17.0 45
ZSM-5 (150)b 315 0.3 2 76 64 5 49 37.9 45
ZSM-5 (500)b 315 0.3 2 39 44 4 17 13.4 45
HZSM-5 without Na 320 1.2 10 100 60 5 60 0.3 63
a
T react = reaction temperature; mcat = mass of catalyst; TOS = time on stream; C = conversion of glycerol; Sacrolein = selectivity for acrolein; Sacetol =
selectivity for acetol; Y acrolein = yield of acrolein; STY = space-time yield; nc = not communicated. b Si/Al ratio.

with a 3D porous network.60 In addition to aluminosilicates, they group, consisting of the replacement of the Brønsted protons
also prepared gallosilicates and ferrosilicates. Nevertheless, good at the catalyst surface by sodium cations through an ion-
catalytic performances were only obtained over the aluminosili- exchange step.63 They used a commercial HZSM-5 catalyst with
cates and gallosilicates, whereas ferrosilicates were less selective a silica/alumina ratio of 65 and varied the number of surface
due to their higher redox-character. The aluminosilicates offered protons between 0.5 and 1. Unsurprisingly, the sodium-free
the most stable performances without any ion exchange, with a catalyst exhibited the best performance in glycerol dehydration,
selectivity for acrolein around 65%. In this patent, the results with a selectivity for acrolein of 60% at full conversion (Table
claimed over zeolitic catalysts were quite similar to those 3), which was explained by a higher number of acid sites, which
reported in Dubois et al.’s work (i.e., 70% selectivity for acrolein were not replaced by sodium.
at full glycerol conversion).39 In further publications, the same Further, Zhou and co-workers proposed a synthesis of micro-
Japanese team reported the tuning of the acidity by modifying and mesoporous ZSM-5 composites using dual templates,
the Si/Al ratio of the zeolite. The best results were achieved subsequently used as catalysts in the dehydration of glycerol
using a ratio of 28, which resulted in yield of acrolein of 63% to acrolein.64 The best result achieved with these catalysts was
(Table 3).61,62 73.6% acrolein selectivity at 98.3% glycerol conversion, which is
Kim et al. studied the influence of the silica/alumina ratio in quite a good performance. The influence of the textural parame-
zeolites of the HZSM-5 type.45 They chose commercial HZSM- ters is not clear for the zeolite compounds used, but the authors
5 catalysts with Si/Al atomic ratios ranging from 23 to 1000. claim that mesopores are more favorable than micropores, as the
A detailed characterisation of the acid properties using am- latter are likely to induce diffusional limitations. A more detailed
monia temperature-programmed desorption revealed an inverse analysis of the impact of the pore size of ZSM-5 zeolites on the
relationship between the total acidity and the silica/alumina selectivity was performed by Pathak et al.65 They showed that the
ratio. In fact, the catalyst with a silica/alumina ratio of 30 selectivity for acrolein increases with the pore size, whereas the
had ca. 9 times more acid sites (66 mmol NH3 g-1 ) than the selectivities to acetaldehyde, formaldehyde and acetol decrease
catalyst with a ratio of 500 (7 mmol NH3 g-1 ). According to the in the same time. Nevertheless, an optimum pore size could not
observed desorption temperature, Kim et al. also showed that be reliably determined, as only four different pore sizes, namely
low silica/alumina ratios result in a higher number of strong acid 0.54, 0.74, 3.15 and 11.2 nm, were investigated.
sites. Nevertheless, these very strong acid catalysts showed only The use of zeolites as catalysts for the dehydration of glycerol
poor selectivity for acrolein (no more than 41.6% for the catalyst offers another attractive means of accomplishing this reaction,
with a silica/alumina ratio of 30, for instance). Therefore, Kim linked with the possibility of tailoring the acid properties
et al. investigated the type of the acid sites and found that by tuning the bulk composition (silica/alumina ratio), by
these low silica/alumina ratios lead to Lewis acid catalysts, incorporation of transition-metal oxides (like gallium oxide or
whereas higher ratios give Brønsted acid catalysts, as already iron oxide) or by modification of the surface (ion-exchange).
postulated by Alhanash et al.44 They concluded that a Brønsted Furthermore, the structure and the mean pore diameter of the
acid catalyst with a high number of sites was more effective for porous network of zeolites can be tuned, whereby – as in the
obtaining a high acrolein yield. These two favorable conditions case of supported inorganic acids – the diffusion limitations can
were gathered in a HZSM-5 with a silica/alumina ratio of 150, be avoided.
which gave 63.8% selectivity at 75.8% conversion at 315 ◦ C
(Table 3). Furthermore, the influence of the reaction temperature
6 Mixed metal oxides, phosphates and
on the last mentioned catalyst was studied, showing that a
reaction temperature below 315 ◦ C gave only low conversion (no
pyrophosphates
more than 50.3%), whereas reaction temperatures above 315 ◦ C As a final group of catalysts used for the dehydration of glycerol,
resulted in poor acrolein selectivity (no more than 57.4%). we summarise in this section the results of various studies
The idea of tuning the number of Brønsted acid sites by concerning metal oxides, phosphates and pyrophosphates.
adjusting the silica/alumina ratio is one possible means of In addition to the aforementioned niobium oxide,40 the group
optimisation. An alternative was proposed by Schüth’s research of Chai studied binary mixtures of zinc oxide, tin oxide, zirconia,

2086 | Green Chem., 2010, 12, 2079–2098 This journal is © The Royal Society of Chemistry 2010
View Article Online

Table 4 Catalytic performances observed over metal oxides, phosphates and oxophosphates, with and without co-feedinga

Catalyst T react (◦ C) mcat (g) Co-feed TOS (h) C (%) Sacrolein (%) Sacetol (%) Y acrolein (%) STY (mmol g-1 h-1 ) Ref.

Nb2 O5 -350 315 0.56 — 10 75 47 10 35 1.3 40


Nb2 O5 -400 315 0.57 — 10 88 51 12 45 1.6 40
Nb2 O5 -500 315 0.61 — 10 91 35 14 32 1.2 40
TiAl-600 315 0.6 — 2 86 43 17 37 1.1 66
Published on 12 November 2010. Downloaded by Pennsylvania State University on 17/09/2016 04:41:30.

SAPO-11 280 0.2 — 1 66 62 10 41 0.9 67


SAPO-34 280 0.2 — 1 59 72 7 42 0.9 67
VPO-700 300 0.2 O2 10 81 95 9 77 6.9 68
VPO-800 300 0.2 O2 10 100 80 6 80 7.1 68
VPO-900 300 0.2 O2 10 97 58 15 56 5.0 68
Fex (PO4 )y 280 0.8 — 5 100 92 0 92 4.1 72
9 wt% WO3 /ZrO2 300 17 — 5 100 74 11 74 1.9 38
19 wt% WO3 /ZrO2 280 5 — nc 83 69 10 57 9.3 79
a
T react = reaction temperature; mcat = mass of catalyst; co-feed = nature of the co-feeding gas; TOS = time on stream; C = conversion of glycerol;
Sacrolein = selectivity for acrolein; Sacetol = selectivity for acetol; Y acrolein = yield of acrolein; STY = space-time yield; nc = not communicated; Mx Oy -***
indicates the calcination temperature (in ◦ C); TiAl = TiO2 /Al2 O3 binary mixture; SAPO = silica-alumina phosphate; VPO = vanadium oxophosphate.

titania, alumina and silica.66 Whereas the catalytic performances pounds increases with the calcination temperature, whereby the
of these mixed oxides remained rather low, with acrolein yields optimum in the catalytic performance (80% yield of acrolein)
not exceeding 36%, the study of the acidity and of the textural was found for a calcination temperature of 500 ◦ C.
parameters clearly showed that the less acid catalysts have The use of phosphates instead of pyrophosphates was studied
increased selectivity for acetol, and that small pore diameters by Deleplanque et al.72 They prepared iron phosphates using
induce a decrease in the selectivity for acrolein. different preparation techniques, such as hydrothermal reaction
Suprun et al. published a paper on phosphate-modified or precipitation. The highest acrolein yield without oxygen co-
titania, alumina and silica/alumina (SAPO),67 and found a feeding was 92.1% (Table 4). In the so-called ‘oxy-dehydration’
strong influence of the acidic and textural parameters on the reaction, i.e., with co-feeding of molecular oxygen, the selectivity
catalytic performances. As in the case of supported inorganic for acrolein decreased to 62.5% at full conversion, due to the total
acids and of zeolites, a relationship between the total acidity and oxidation to carbon oxide species (COx ) as an important side
the selectivity for acrolein was observed. In fact, the most acidic reaction. Therefore, Dubois investigated the addition of group 1
catalyst, a silica-alumina phosphate, showed a selectivity for and group 2 metal cations to these catalytic formulations, such
acrolein up to 72% (Table 4). Furthermore, Suprun et al. pointed as caesium, strontium and potassium.73 Doping with caesium in-
out that microporous materials (5–6 Å) are less active due creased the acrolein yield to 72.2%, which was thus significantly
to internal diffusion limitations, and tend to induce increased higher than that observed over the undoped iron phosphate
carbon deposition, as observed on SAPO-34. (62.5%), but still lower than the yield obtained without oxygen
A similar catalytic system, namely vanadium oxophosphates co-feeding (92.1%). The idea of doping the phosphate species
and oxo-pyrophosphates, was studied by Wang et al.68,69 As was also investigated by Matsunami et al., who used phosphate-
in the case of niobium oxide, the acidity of these compounds modified silica with alkali salts. Nevertheless, in this latter case,
can be tuned by modifying their calcination temperature, but the yield of acrolein did not exceed 67%.74,75
in an inverse relationship, as high calcination temperatures In 2006, Dubois et al. proposed the use of tungsten oxide on
result in a larger number of acid sites. The catalysts were zirconia as a catalyst for glycerol dehydration. Total conversion
tested in the glycerol dehydration reaction, with co-feeding of glycerol was achieved at 300 ◦ C with a catalyst containing
of oxygen. The best result was obtained using a vanadium 9 wt% of tungsten oxide, with a yield of acrolein of 74%.38
pyrophosphate calcined at 800 ◦ C, which gave a selectivity Redlingshofer et al. further studied this binary metal oxide
for acrolein of 66% at full conversion (Table 4). Nevertheless, system.76,77 The reaction was carried out at a lower temperature
this result was not obtained over the catalyst with the largest (260 ◦ C). Good performances were observed with an acrolein
number of acid sites. Actually, as the specific surface decreased yield between 77 and 79%. Nevertheless, the catalysts tended
for high calcination temperatures due to sintering of the solid, to deactivate on-stream, with the acrolein yield decreasing at a
the selectivity consequently dropped drastically in this case. rate of 5% every 10 h. Additionally, these authors described
Dubois extended the study of pyrophosphates using boron and the possibility of regenerating the catalysts by oxygen post-
aluminium as cations, and observed that the boron-containing treatment at 350 ◦ C for 5 h. Bythis treatment, they claimed
compounds give the highest acrolein yields, with a maximum of that the initial catalytic performances could be recovered.78
77.8%.70 This well-known combination of tungsten oxide on zirconia
Other metal pyrophosphates were studied by Liu and cowork- was further studied by Ulgen et al.79 They showed that the acidity
ers, who used rare earth metals such as lanthanum, neodymium of the solid increases with the amount of tungsten oxide. For the
and cerium.71 All the considered pyrophosphates exhibited catalyst containing 19 wt% of tungsten oxide, an acrolein yield
quite similar catalytic performances, with the best results being of 57% was observed (Table 4).
achieved over Nd4 (P2 O7 )3 . As in the case of the aforementioned Furthermore, they studied the influence of the calcination
vanadium pyrophosphates, the acidity of the neodymium com- temperature on the physical properties of the catalysts. Whereas

This journal is © The Royal Society of Chemistry 2010 Green Chem., 2010, 12, 2079–2098 | 2087
View Article Online

the temperature treatment had only a low impact on the acidic Furthermore, the hydrogen-to-carbon (H/C) ratio for the
properties, the number of basic sites significantly decreased when carbon species was determined by elementary analysis. All the
the calcination temperature increased. Moreover, the textural catalysts showed a decrease in the H/C ratio at higher reaction
parameters were quite dramatically modified, as the sintering temperatures, meaning that the carbon deposit became more
led to larger pores but reduced specific surface areas. For these deficient in hydrogen (Table 5). For example, the H/C ratio over
two reasons, the thermal treatment had in any case a positive SAPO-34 decreased from 0.62 at 280 ◦ C to 0.47 at 320 ◦ C.
impact on the catalytic performance, as the decrease in basicity This can be ascribed to the formation of highly condensed,
Published on 12 November 2010. Downloaded by Pennsylvania State University on 17/09/2016 04:41:30.

led to the suppression of acetol formation, and the larger pore unsaturated carbonaceous aromatic species. Solutions still have
size reduced the deactivation extent due to coking, as internal to be found to avoid or to limit the coke deposit on the catalyst
diffusion limitations were lower. surface or at least to optimise the regeneration of the catalysts.
The above-described family of mixed metal oxides, phosphates Although a fundamental approach is still needed for studying
and pyrophosphates is rather heterogeneous. Some compounds, efficient means for limiting the coke deposition during the
like niobium oxide, tungsten oxide and pyrophosphates, offer the reaction, at least three kinds of solutions have been proposed so
possibility of controlling their acid strength via the calcination far for continuously regenerating the catalysts: (i) co-injection
step. This thermal treatment generally also has an influence on (co-feeding) of oxygen or hydrogen with the gas feed to yield
the specific surface area and the pore size due to sintering. in situ regeneration;39 (ii) cyclic regeneration of the used catalyst
In comparison to the aforementioned zeolites and supported by injection of a flow or of pulses of air or oxygen;83 and
HPAs, the control of the physical properties of the mixed metal (iii) circulation of the catalyst in a moving-bed reactor with
oxides and phosphates seems nevertheless less straightforward. regeneration in a separate parallel vessel (as for the FCC
process).25 Whereas the first option is associated with the risk of
generating explosive conditions and/or to oxidise the various
reaction products, the second alternative is accompanied by
7 Regeneration of spent catalyst the disadvantage of a loss in productivity. The third option
From the studies reported above, it is clear that very efficient does not suffer from these drawbacks, but the construction
catalysts for the dehydration of glycerol to acrolein can be and the operation of a circulating bed reactor implies serious
prepared. Unfortunately, these catalysts quickly deactivate, most technological difficulties that can only be overtaken by skilled
probably because of the formation of coke, which makes it persons through fine process control.
difficult to straightforwardly use them in industrial plants at The idea of working with air in the reactant mixture was first
the current time. Up to now, only a few studies have been proposed by Dubois for catalysts of the zeolite type.39 To avoid
devoted to the characterisation of carbon deposits and to the the explosivity range, the oxygen fraction must never overpass
sites responsible in these catalytic systems. Erfle et al. used FTIR 7 vol%. Therefore, the composition of the reaction feed was
spectroscopy to characterise the spent catalyst and observed adjusted to 6% of oxygen for 4.5% of glycerol, the remaining
C O and C C signals.56 They attributed these signals to cyclic 89.5% being steam. Thereby, the authors claim to reduce catalyst
anhydride species and claimed that the Brønsted acid sites were deactivation and even to inhibit the formation of by-products
the centers of coke formation. Suprun and coworkers confirmed such as acetol in these conditions.
that the acid sites were responsible for the formation of carbon The same process was further used for vanadium
deposits on phosphate catalysts.67 They further found that the pyrophosphates,69 boron phosphates,70 iron phosphates,72 tung-
quantity of coke increased with higher reaction temperatures sten oxide on zirconia79 and caesium salts of phosphotungstic
and also with decreasing the pore diameter of the catalyst (Table acid.57 As a general trend, the selectivity for by-products such
5). For example, the coke loading on SAPO-11 increased from as acetol, acetaldehyde and formaldehyde decreased, whereas
4.6 wt% at 280 ◦ C to 9.5 wt% at 320 ◦ C. From the ammonia the formation of organic acids, like acrylic acid, acetic acid
desorption experiments, one can also see that the acid strength and formic acid, increased, due to subsequent oxidation of the
increases in the order Al2 O3 -PO4 ª TiO2 -PO4 < SAPO-11  aldehydes formed. For some catalysts, like zeolites, vanadium
SAPO-34, which correlates well with the amount of carbon oxophosphates, and tungsten oxide on zirconia, the selectivity
deposit found over these catalysts. Indeed, the quantity of for acrolein was not or only slightly affected. In the case of
deposited carbon increased for all the reaction temperatures the caesium salt of phosphotungstic acid, an increase in the
in the same order. selectivity for acrolein by a factor of two was even observed (93%

Table 5 Correlation between acidity and the carbon deposit for phosphate catalystsa

T react = 280 ◦ C T react = 300 ◦ C T react = 320 ◦ C

Coke loading H/C Coke loading H/C Coke loading H/C


Catalyst Ac (mmol NH3 g-1 ) Dp (Å) SBET (m2 g-1 ) (wt%) ratio (wt%) ratio (wt%) ratio

Al2 O3 -PO4 295 111 118 2.4 0.54 3.9 0.52 5.8 0.49
TiO2 -PO4 258 101 38 3.1 0.55 6.8 0.50 8.9 0.48
SAPO-11 1330 6 172 4.6 0.68 7.4 0.55 9.5 0.40
SAPO-34 498 5 49 9.6 0.62 12.7 0.54 16.2 0.47
a
Ac = total acidity; Dp = pore diameter; SBET = specific surface area; T react = reaction temperature; SAPO = silica-alumina-phosphate.

2088 | Green Chem., 2010, 12, 2079–2098 This journal is © The Royal Society of Chemistry 2010
View Article Online

Table 6 Comparison of catalytic performance observed in glycerol dehydration, with and without co-feedinga

Catalyst T react (◦ C) mcat (g) Co-feed TOS (h) C (%) Sacrolein (%) Sacetol (%) Y acrolein (%) STY (mmol g-1 h-1 ) Ref.

b-Zeolite 300 4.2 O2 3 100 57 0 57 5.8 39


b-Zeolite 300 4.2 — 1.5 100 57 10 57 5.8 39
HZSM-5 300 6.4 O2 1 88 46 2 40 2.7 39
HZSM-5 300 6.4 — 1 79 49 6 39 2.6 39
Published on 12 November 2010. Downloaded by Pennsylvania State University on 17/09/2016 04:41:30.

VPO-800 300 0.2 O2 nc 100 65 6 65 5.8 68


VPO-800 300 0.2 — nc 95 60 12 57 5.1 68
BPO 280 0.3 O2 nc 100 65 8 65 0.8 70
BPO 280 0.3 — nc 96 81 6 78 0.9 70
Fex (PO4 )y 280 1.3 O2 5 100 63 0 63 1.7 72
Fex (PO4 )y 280 0.8 — 5 100 92 0 92 4.1 72
19 wt% WO3 /ZrO3 280 5 O2 nc 73 74 8 54 8.9 79
19 wt% WO3 /ZrO2 280 5 — nc 83 69 10 57 9.3 79
CsPW 260 23 O2 nc 100 93 nc 93 3.0 57
CsPW nc 23 — 1 83 47 3 39 0.7 57
0.5% Pd-CsPW 275 0.3 H2 5 79 96 1 76 37.9 44
CsPW 275 0.3 — 1 100 98 nc 98 49.0 44
WO3 /ZrO2 275 9.2 SO2 24 87 69 7 60 1.2 81
WO3 /ZrO2 275 9.2 — 24 69 71 9 49 1.0 81
a
T react = reaction temperature; mcat = mass of catalyst; Co-feed = nature of the eventual co-feeding gas; TOS = time on stream; C = conversion of
glycerol; Sacrolein = selectivity for acrolein; Sacetol = selectivity for acetol; Y acrolein = yield of acrolein; STY = space-time-yield; nc = not communicated;
VPO = vanadium oxophosphate calcined at 800 ◦ C; BPO = boron phosphate; CsPW = Cs2.5 H0.5 PW12 O40.

vs. 47%; Table 6). On the other hand, for boron phosphates and significantly inhibited the formation of acetol (selectivity of
iron phosphates, the co-feeding of oxygen led to a significant 0.2% vs. 2.4% without SO2 ). After 24 h on-stream, the catalytic
decrease in the selectivity for acrolein, which was ascribed to performance observed when co-feeding SO2 were significantly
the over-oxidation to carbon oxides (CO/CO2 ). To facilitate the improved, still with an excellent conversion of 87% (initially
activation of molecular oxygen, doping with metal cations like 100%), whereas it was 69% in the absence of SO2 (Table 6).
caesium, potassium, strontium, silver and platinum was probed The results obtained using co-feeding of oxygen are quite
by Dubois, which led to an acrolein selectivity of around 70%.73 different and strongly depend on the studied catalytic system. As
A comparable approach was followed by Kasuga et al., who mentioned earlier, the oxygen concentration must be kept below
used protonated MFIs as catalysts.80 When they injected air in 7 mol% in order to stay out of the explosion limits. Wang et al.
the gas reactant feed, the selectivity for acrolein on unmodified studied the influence of the oxygen concentration on the catalytic
catalysts was rather low and did not exceed 45% after 24 h performance of vanadium phosphate catalysts. They observed
under reactant flow. To optimise the regeneration, Kasuga et al. that too little (as well as too much) oxygen has a negative impact
thus also proposed to modify the MFI protonated-zeolite with on the selectivity for acrolein.69 In fact, low quantities of oxygen
a small amount of metal (Pt, Pd, Ru, Cu, Ir or Au). By this led to larger amounts of acetol, whereas high concentrations
mean of doping, they claimed that it is possible to accelerate the induced an increase in the selectivity for acetaldehyde, acetic
splitting of the dioxygen used during regeneration.80 The best acid, acrylic acid and – of course – to carbon oxides, which
results were obtained with 0.1 wt% Pt and 1 wt% Au (80.7 and is explained by the successive over-oxidation of the products.
79.7% of acrolein yield, respectively, at full glycerol conversion The exceptions where the co-feeding process is unfavorable to
after 150 min under reaction conditions). the selectivity for acrolein – as over iron phosphate and boron
The same concept to reduce the carbon deposit, but this time phosphate – can be easily explained by the redox character of
by co-feeding hydrogen, was investigated by Alhanash et al., the catalysts, which promotes oxidation processes. In these cases,
who used caesium salts of HPAs.44 Following a similar principle co-feeding with hydrogen might be a good way to overcome this
as that proposed by Kasuga et al.,80 they probed the doping with problem. The process using co-feeding of sulfur dioxide might
small quantities of noble metals such as ruthenium, palladium also be an interesting approach, even if one should keep in mind
and platinum. Thereby, they could increase the catalyst activity its toxicity, which is certainly the most important drawback of
by a factor of two after 5 h of reaction (conversion of 79% vs. this method. Furthermore, this reagent may be incompatible
41%; Table 6) without any significant impact on the selectivity with some catalytic systems, due to some possible poisoning
for acrolein (96% vs. 98%). The best results were obtained with issues.
palladium on the caesium salt of phosphotungstic acid, over The second process for regenerating the spent catalyst –
which a productivity of 37.9 mmol g-1 h-1 was achieved. namely, periodic regeneration – was studied by Arita et al., who
Dubois proposed a third kind of feed additive for enhancing regenerated a used H-ZSM5 catalyst under an air flow. They
the long-lasting performances of tungsten oxide on zirconia. evidenced the possibility of recovering the initial performances.82
Instead of oxygen or hydrogen, he injected 250 ppm of sulfur However, the exothermicity of the regeneration led to the
dioxide into the feed in order to slow down the deactivation.81 In formation of hot-spots. Consequently, the temperature during
fact, the addition of SO2 even induced a slight positive impact regeneration exceeded the reaction temperature by more than
on the selectivity for acrolein (73% vs. 72% without SO2 ) and 100 ◦ C, at which the catalyst risks thermal decomposition.

This journal is © The Royal Society of Chemistry 2010 Green Chem., 2010, 12, 2079–2098 | 2089
View Article Online

Therefore, this technique is not applicable for temperature- pretreatments induced higher acrolein selectivity in the first
sensitive catalytic systems. few minutes after switching to the reactant flow. However,
A more academic approach was followed by Atia et al. after 150 min on-stream, the catalytic performances returned
Their preliminary work involved long-term runs (up to 300 h), to values similar to those obtained without any preliminary
and they found that over alumina-supported heteropolyacid pretreatment, which means that this effect is only effective in the
catalysts the selectivity for acrolein remained quite stable, early stages of the reaction. Despite this, the authors pointed
whereas glycerol conversion more-or-less linearly decreased. out the possibility of recovering the initial performance by
Published on 12 November 2010. Downloaded by Pennsylvania State University on 17/09/2016 04:41:30.

This effect is typically observed for deactivation by deposition regenerating the catalysts at 500 ◦ C under a flow of air.
of coke on the catalyst surface, and is linked with the progressive
decrease in the number of accessible active sites. To verify this
hypothesis, spent catalysts were regenerated under a flow of 1%
8 Dehydration of glycerol in the liquid phase
of oxygen in nitrogen at 325 ◦ C for 24 h. After this treatment, the As mentioned in the introduction, the first experimental results
regenerated catalyst showed performances identical to those of on dehydration of glycerol in the liquid phase were reported
the fresh catalyst. To quantify the coking effect, the spent catalyst by Groll in 1936,33 and by Hoyt and Manninen in 1951,34 in the
was also analysed by thermogravimetry, which straightforwardly presence of either homogenous or heterogeneous catalysts. Since
confirmed the hypothesis of carbon deposition.52 these studies, less than a dozen articles and patents have been
Meanwhile, Corma et al. adopted the idea of a circulating published on liquid-phase dehydration, while there has been a
bed reactor,25,83 which was previously disclosed in a Dubois boom in the number of publications dealing with the gas-phase
patent.38,39 These authors studied the opportunity of injecting reaction. In the following section, we give a brief overview of
crude glycerol directly into FCC plants. An advantage would the dehydration reaction in the liquid phase, and point out the
be the use of existing facilities and therefore avoiding having major drawbacks of this process.
to invest in building specific infrastructures. Note that this idea Ramayya et al. studied the dehydration of glycerol in batch
was already proposed, but under another form, by Dubois, who experiments using an aqueous solution of sulfuric acid, and
pointed out the possibility of injecting glycerol into propylene conditions close to the critical point of water (namely P = 22.1
oxidation plants to concomitantly yield acrolein from both MPa, and T = 374 ◦ C).86 (Near) supercritical water offers the
sources.84 In the FCC plants, the heat recovered by the burning advantage that its physical properties, like the dielectric constant
of coke could be used to provide the energy necessary for or the ion product, can be adjusted by varying temperature and
the evaporation of glycerol. Corma et al. concluded that an pressure. The mixture was heated in a batch reactor at 350 ◦ C,
autothermal process is possible in that case. Nevertheless, the pressurised to 34.5 MPa and acidified with 5 mM sulfuric acid.
authors did not limit their view to the dehydration reaction, but In these conditions, and at a residence time of 25 s, the conversion
also investigated the possibilities for reforming glycerol at 500– of glycerol and the selectivity for acrolein were 55% and 86%,
600 ◦ C to produce ethylene and propylene. In this approach, respectively (Table 7). Considering the fact that sulfuric acid is
the glycerol feedstock may be an alternative or a complement to soluble in supercritical water, this process can be classified as
naphtha cracking. homogeneous catalysis with, the consequent catalyst–reaction
An original approach was followed by Kasuga et al., who mixture separation issues. Furthermore, a batch process is not
focused their investigations on various possibilities for pre- applicable for large-scale production of acrolein.
treating catalysts (protonated MFI zeolite or Alox alumina).85 Nevertheless, the idea of using supercritical water as a reaction
Three types of pretreatments were tested: (i) a flow of acetol, medium was a new aspect, and Bühler et al. decided to study it
water and nitrogen; (ii) a flow of acrolein, water and nitrogen; in more detail in 2001.87 Instead of working in a batch reactor,
and (iii) a flow of acetol, water and air. Irrespective of the they built a flow-type reactor and used supercritical water both
pretreatment method, the Alox catalyst always led to very as a solvent and a catalytic system. They observed various
poor acrolein yields, while over the MFI catalyst all these decomposition products such as acetaldehyde, formaldehyde,

Table 7 Catalytic performances observed for glycerol dehydration in the liquid phasea

Catalyst T react (◦ C) P (MPa) Solvent Reactor type C (%) S (%) Y (%) Ref.

H2 SO4 190 0.1 Water Batch nc nc 49 33


H3 PO3 /clay 300 0.1 Paraffin Batch nc nc 72 34
H2 SO4 350 34.5 Water Batch 55 86 47 86
No catalyst 356 45 Water Continuous 71 38 27 87
ZnSO4 360 25 Water Continuous 50 75 38 88
ZnSO4 360 34 Water Continuous 62 59 37 89
H2 SO4 400 34.5 Water Continuous 92 81 74 90
KHSO4 280 0.1 Paraffin Continuous 97 82 80 91
MgSO4 280 0.1 Paraffin Continuous 92 56 52 91
5 wt% H3 PO4 /Al2 O3 280 0.1 Paraffin Continuous 89 57 51 91
FePO4 320 0.1 Xylene Continuous 87 78 68 93
CuPO4 280 0.1 Sulfolane Continuous 98 85 84 93
a
T react = reaction temperature; P = pressure; C = conversion of glycerol; S = selectivity for acrolein; Y = yield of acrolein; nc = not communicated.

2090 | Green Chem., 2010, 12, 2079–2098 This journal is © The Royal Society of Chemistry 2010
View Article Online

allylic alcohol and propionaldehyde, as well as acrolein. The Thus, we have seen that the dehydration of glycerol in the
acrolein yield varied between 10 and 27% depending on the liquid phase (at atmospheric pressure or in near-critical or
reaction conditions, with the best result (27%) obtained at a supercritical conditions) leads to the formation of acrolein. The
reaction temperature of 350 ◦ C under 45 MPa pressure (Table 7). application of supercritical conditions is an especially interesting
As mentioned earlier, these tests were carried out in the absence concept – the change in the physical properties of water inducing
of any catalyst or acid component, which explains the poor yields dehydration of glycerol even in the absence of a catalyst –
in acrolein. but, unfortunately, the yield of acrolein remains rather low in
Published on 12 November 2010. Downloaded by Pennsylvania State University on 17/09/2016 04:41:30.

Recently, a combination of the ideas of Hoyt and Manninen34 this case. On the other hand, the combination of acids and
and of Ramaya et al.,86 namely using a supported catalyst in near-critical or supercritical conditions results in much better
supercritical water as a reaction medium, was proposed by Ott performances, but also in the generation of an extremely corro-
and coworkers.88 They investigated the dehydration of glycerol in sive medium. Therefore, the reaction vessel, which must resist
a batch reactor under supercritical conditions using zinc sulfate high temperature and pressure, must be specifically designed,
as a catalyst. The choice of zinc sulfate as a catalyst is quite implying important investment/maintenance costs, which is a
surprising, as this compound is not known as a strong acid major drawback for industrial applications.
catalyst, which is typically needed for this reaction. However, Recent results show that the initial idea of using high-boiling
they justified their choice considering material stress issues: liquids as reaction media at atmospheric pressure seems more
water in its supercritical form is very corrosive, and therefore preferable. Furthermore, the proposed process in this latter case,
requires the use of special and expensive steel grades for reactors; where the reactant is constantly fed into the reaction mixture
this corrosive power would become even stronger if acidic and the acrolein is separated by evaporation upon formation,
compounds were subsequently added to the medium, leading can be classified as a continuous process. Nevertheless, the
to unacceptable reactor material stress. In these conditions, the accumulation of heavier by-products in the reaction mixture
authors showed that supercritical conditions are not necessarily seems quite probable in this case. Moreover, the adaptation of
the optimal ones. Actually, the best catalytic performance, this kind of process for industrial application is a real challenge,
namely 75% selectivity at 50% conversion, was obtained in the as the results from laboratory scale (300 mL) cannot easily be
near-critical region of water (310 ◦ C at 25 MPa) (Table 7). This transferred to plant size. In conclusion, one can state that a more
might be explained by the degradation of acrolein when it is practical process seems to still be a topical issue of research
subjected to temperatures that are too high. In 2007, Lehr et al. before commercial applications of liquid-phase dehydration of
also published results on the dehydration of biomass-derived glycerol to acrolein can be reliably envisaged.
polyols in sub- and super-critical water.89 As far as glycerol is
concerned, they reported 59% glycerol conversion with almost
9 Reaction mechanism
60% selectivity for acrolein over zinc sulfate as a catalyst under
subcritical water conditions, like Ott et al. (Table 7).88 Whereas the research for an efficient catalyst can follow (at least
To determine if the dehydration of glycerol in the liquid phase in an initial screening step) a more-or-less applied approach,
might offer a sustainable source of acrolein at an industrial scale, a more fundamental aspect lies in the understanding of the
Watanabe et al. attempted to optimise the acrolein yield.90 They reaction mechanism, which further enables back-optimisation
combined the supercritical conditions in a flow-type reactor with of the catalytic formulations. This requires the identification of
sulfuric acid as a catalyst. Several blank tests performed in the the intermediate steps and the explanation of the formation of
absence of a catalyst gave low glycerol conversions, but high the by-products. A first proposal was made by Bühler et al. for
acrolein selectivities with a yield up to 16%. The best results the glycerol activation in the liquid phase.87 Two pathways were
were obtained at 400 ◦ C under 34.5 MPa with a sulfuric acid claimed, either via an ionic or a radical mechanism (Scheme
concentration of 5 mM. In these conditions, the yield of acrolein 8). The ionic reaction (Scheme 8a) begins with the protonation
was 74% at a glycerol conversion around 92% (Table 7). One of of glycerol either on one primary hydroxyl group or on the
the main issues in this work was to increase the formation rate secondary hydroxyl group. Consequently, the elimination of a
of acrolein without modifying its decomposition rate, which was water molecule leads to the formation of a carbocation. In the
achieved by adding acid in supercritical conditions. case of the secondary carbocation, the only possible product
In 2007, Suzuki et al. and Yoshimi et al. revamped the idea after release of H3 O+ is acrolein, whereas either acrolein or
of Hoyt and Manninen,34 using a liquid reaction medium with acetaldehyde + formaldehyde can be formed starting from the
high boiling point at atmospheric pressure.91,92 Potassium sulfate primary carbocation. The radical pathway (Scheme 8b) starts
and bisulfate were used as catalysts in a batch reactor, where with the abstraction of a hydrogen atom from a primary carbon
the glycerol solution was added dropwise in a paraffin solution using an OH∑ radical. The resulting radical species further loses
maintained at 280 ◦ C. The produced acrolein was continuously an hydroxyl radical (OH∑ ), which leads to the formation of
evaporated from the reaction mixture and was recovered in an acrolein.
aqueous solution, and yields close to 80% were achieved. This Tsukuda et al. – who performed the dehydration in the gaseous
technique was further adapted by Takanori et al., who used iron phase – developed a more formal reaction framework (Scheme
or copper phosphate as catalysts in high-boiling organic solvents 9).51 The first dehydration step leads to the formation of two
(e.g. n-octane, m-xylene or sulfolane) as reaction media.93,94 They enols, which are in tautomeric equilibrium with the correspond-
obtained (to our knowledge) the highest ever reported acrolein ing ketone (acetol) and aldehyde (3-hydroxypropionaldehyde).
yield in the liquid phase, over copper sulfate in sulfolane (84%; The latter may subsequently react in a second dehydration
Table 7). reaction to yield acrolein, or undergoes a retroaldol reaction to

This journal is © The Royal Society of Chemistry 2010 Green Chem., 2010, 12, 2079–2098 | 2091
Published on 12 November 2010. Downloaded by Pennsylvania State University on 17/09/2016 04:41:30. View Article Online

Scheme 8 Reaction pathways to the formation of acrolein considering


(a) an ionic pathway and (b) a radical pathway.87

liberate formaldehyde + acetaldehyde, which – in the presence of


oxygen – can easily be oxidised to acetic acid. In agreement with
the previous proposal of Bühler et al.,87 Tsukuda et al. confirmed
that the key to obtain high acrolein selectivity lies in the control Scheme 10 Reaction network proposed by Corma et al.25
of the first dehydration step, whereby the formation of 3-
hydroxypropionaldehyde must be favored over the formation of
acetol, which is identified as the main by-product of the process.
The same mechanism has been further amended by Chai a reactant at 350 ◦ C. The products were classified into three
et al. (Scheme 9),41 who added two possible hydrogenation groups: acids (9.4%), other aldehydes (52.0%) and coke (27.7%).
steps, one starting from acrolein and the other from acetol Furthermore, the formation of low amounts of acetone (4.0%),
to yield allylic alcohol and 1,2-propanediol, respectively. In acetaldehyde (1.4%) and carbon monoxide (1.2%) was observed.
the catalytic system studied, these by-products can indeed be Nevertheless, the conversion of acetol did not exceed 25%, which
observed. Furthermore, the hydrogenation of formaldehyde to means that this compound is quite stable and does not easily
methanol or its decomposition to CO + H2 are mentioned. undergo consecutive reactions. This explains why it is usually
identified as the major by-product in the reaction of dehydration
of glycerol to acrolein.
Even though only small amounts of acetone were observed,
Corma et al. investigated the consecutive reactions starting from
this compound by repeating their method. The resulting product
distribution was dominated by unsaturated hydrocarbons like
butene (27.9%), propylene (1.7%), C5 + C6–8 aromatics (6.6%)
and – as a product of these compounds – coke (31.9%). Other
identified products were various acids (20.2%) and aldehydes
(7.3%), but, as in the case of hydroxylacetone, the conversion
of acetone remained low and did not exceed 14%. Nevertheless,
with their findings, they enlarged the reaction scheme proposed
by Chai and Tsukuda by adding further side-reactions ema-
nating from acetol. Thereby, they explained the formation of
oligomers and coke by consecutive reactions starting from acetol
via acetone and acetaldehyde.
Scheme 9 Mechanism proposed by Tsukuda et al. and further com- The idea of investigating the consecutive reactions starting
pleted by Chai et al.41,51 from a by-product was picked up by Suprun et al. for 3-
hydroxypropionaldehyde and acetol.67 They could thus confirm
These two first proposals were the first approaches for the retroaldol reaction previously postulated by Tsukuda,51
explaining the observed by-products from glycerol dehydration. which leads to the formation of formaldehyde and acetaldehyde
Successively, Corma et al.25 postulated a more complex interwo- issuing from 3-hydroxypropionaldehyde. Additionally, signifi-
ven reaction framework (Scheme 10), where the main reaction cant amounts of coke were found, which was ascribed to the
steps are the same as those previously described by Tsukuda and formation of cyclic C6 compounds identified by GC–MS analysis
Chai. In order to achieve a deeper understanding and specifically (Scheme 11a). The authors further identified numerous furan
investigate the secondary reactions, the authors used acetol as derivatives as products from acetol (Scheme 11b).

2092 | Green Chem., 2010, 12, 2079–2098 This journal is © The Royal Society of Chemistry 2010
View Article Online

results in a new type of product distribution. Indeed, new


by-products, such as organic acids, are observed, and a new
reaction scheme has thus been very recently proposed by
Deleplanque et al. (Scheme 12).72 In addition to dehydration,
which yields the already known products like acetol and
acrolein, or their reductions to acetone and allylic alcohol, the
authors postulate some new oxidation reactions. Starting from
Published on 12 November 2010. Downloaded by Pennsylvania State University on 17/09/2016 04:41:30.

formaldehyde, originating from the aforementioned retroaldol


reaction, oxidation leads to the formation of formic acid, which
is easily over-oxidised to CO or CO2 . When acrolein is oxidised,
acrylic acid is observed, which may further react to give acetic
acid by decarbonylation. Alternatively, the latter can also be
formed by oxidation of acetaldehyde. On the other hand, the
formation of propanoic acid is explained by the reduction of
allylic alcohol in a first step and subsequent oxidation of the
formed 1-propanol. In addition to these oxidation reactions, the
authors also describe the formation of acetals from condensation
of glycerol with formaldehyde. This may be possible when the
glycerol concentration is rather high compared to the oxygen
concentration, whereby the oxidation of formaldehyde to formic
acid becomes limited. Nevertheless, one can conclude that
the presence of oxygen results in a much higher number of
by-products, which explains the lower selectivity for acrolein
observed in the case of the dehydration reaction conducted in
the presence of oxygen.
All these reaction mechanisms are mostly formal and actually
do not explain the real interactions with the surface of the solid
catalysts. Therefore, Yoda et al. developed a mechanism for the
dehydration over H-MFI zeolites inspired by the results of in situ
Scheme 11 Products derived from (a) 3-hydroxypropionaldehyde and infrared observations.95 The mechanism is similar to the afore-
(b) acetol.67
mentioned ones, whereby glycerol interacts with the bridging
hydroxyl-group of the zeolite to give alkoxy species (Scheme
In the previous section, we have seen that the co-feeding of 13). When glycerol is bound by a terminal OH-group, acetol
oxygen has a positive impact on the long-term performances is formed by the known keto–enol tautomerism. On the other
of the catalysts. Nevertheless, the presence of oxygen also hand, the interaction with the secondary OH-group leads to

Scheme 12 Reaction scheme proposed for dehydration of glycerol in the presence of oxygen.72

This journal is © The Royal Society of Chemistry 2010 Green Chem., 2010, 12, 2079–2098 | 2093
Published on 12 November 2010. Downloaded by Pennsylvania State University on 17/09/2016 04:41:30. View Article Online

Scheme 13 Reaction mechanisms over (a) Brønsted and (b) Lewis acid sites as proposed by Alhanash et al.95

the formation of an intermediate of 3-hydroxypropionaldehyde, of the secondary hydroxyl-group (Fig. 5). In contrast, the
and, after a second dehydration step, to the formation of second dehydration step – starting from the intermediate 3-
acrolein. hydroxy-propionaldehyde – required only 29.7 kcal mol-1 . The
Based on these proposed reaction mechanisms, two research dehydration of the primary OH-group, leading to the formation
groups calculated the energy states for the different reaction of acetol, was found to be energetically even more unfavorable,
pathways. Nimlos et al. concentrated on the calculations of the with an activation energy of 73.2 kcal mol-1 . Nevertheless, the
neutral and the protonated reaction pathways using the Gaus- formation of acetol is exothermic and liberates 4.0 kcal mol-1 ,
sian 03 algorithm with B3LYP/6-311G sets for the molecular whereas acrolein requires +9.3 kcal mol-1 .
orbitals.96 For the so-called ‘neutral glycerol’, they found a very The calculations for the protonated glycerol – as proposed by
high activation barrier of 70.9 kcal mol-1 for the dehydration Bühler et al.87 – show that the activation energies are much lower

Fig. 5 Calculation for neutral and protonated reaction pathway to acrolein and acetol.96

2094 | Green Chem., 2010, 12, 2079–2098 This journal is © The Royal Society of Chemistry 2010
Published on 12 November 2010. Downloaded by Pennsylvania State University on 17/09/2016 04:41:30. View Article Online

Fig. 6 Calculated energies for the glycerol dehydration over MFI-zeolite.97

than for the non-protonated one. Starting from the protonated 10 Direct use of crude glycerol as a reactant: an
secondary hydroxyl group of glycerol, the activation energy of important economic issue
the transition state of the first dehydration is only 22.4 kcal mol-1
vs. 70.9 kcal mol-1 without proton. The second dehydration As mentioned in the introduction, the price of glycerol drasti-
step, which is calculated to 24.6 kcal mol-1 in the presence cally depends on the required technical grade. Whereas refined
of a proton, is also 5.1 kcal mol-1 lower than without (29.7 glycerol cost between 500 and 600 US$ t-1 in January 2010, crude
kcal mol-1 ). Furthermore, the reaction pathway for protonated glycerol could be obtained at the same time for only 200 US$
glycerol leads to the exothermic formation of acrolein with an t-1 .26 Nevertheless, the glycerol market has previously proven to
energy release of 18.0 kcal mol-1 . As in the case of the non- be highly volatile, with costs up to 1800 US$ t-1 for refined glyc-
protonated reaction, the formation of acetol requires a higher erol and up to 500 US$ t-1 for crude glycerol in 2007–2008.26 It
activation energy of 24.9 kcal mol-1 compared to that required should also be mentioned that glycerol availability and price also
for the formation of acrolein (22.4 kcal mol-1 ). Moreover, the depend on political decisions concerning biodiesel production
differences in the final energy states of the two products are quite and agricultural subsidies. Therefore, making possible the use
low (acetol: -18.3 kcal mol-1 ; acrolein: -18.0 kcal mol-1 ). From of crude glycerol is a crucial issue for achieving a sustainable
these results, one can conclude that the formation of acetol is and economically viable process for the production of acrolein.
thermodynamically less favored than the formation of acrolein, However, as previously mentioned, crude glycerol is generally
due to the associated higher barrier of activation energy. contaminated with by-products issued from biodiesel process
Nevertheless, the difference is rather small, which suggests (inorganic traces like salts and organic traces like water, esters,
that it might be difficult to entirely suppress the formation of fatty acids and alcohols). The use of crude glycerol as a feedstock
acetol. may therefore induce serious difficulties, either by poisoning of
The calculations based on the results obtained for the reaction the catalysts or by causing plugs due to deposition of high-
over ZSM-5 zeolite catalysts– as proposed by Yoda et al.95 – were boiling organic materials or inorganic salts. To our knowledge,
performed by Kongpatpanich et al., who used Gaussion 03 with at this time no catalytic test using this type of feedstock has
M06-2X/6-31G set for the zeolite 12T cluster model.97 For the been reported in the literature – all the disclosed studies were
first dehydration step, they found an activation energy slightly carried out using diluted aqueous solutions prepared from
higher than that obtained by Nimlos et al. for the protonated refined glycerol. Furthermore, it has been found that the yield
glycerol (41.4 kcal mol-1 vs. 22.4 kcal mol-1 (Fig. 6).96 The of acrolein strongly depends on the glycerol concentration in
activation energy of the second dehydration was calculated at the reactor feed, with a decrease in the yield with increasing
38.5 kcal mol-1 and the final energy state was +19.5 kcal mol-1 . concentration.35,36 However, the use of crude glycerol is of prime
It is difficult to explain these differing results and, unfortunately, economical importance, and different ideas have already been
Kongpatpanich has not yet calculated the reaction pathway proposed to avoid the costly distillation of crude glycerol and to
leading to the formation of acetol. overcome the obstacles raised by its use as a feedstock. Kijenski
The mechanisms described give quite good explanations for et al. proposed a modified evaporation system, where the crude
the observed products in the glycerol dehydration reaction, such glycerol is brought into contact with an inert liquid at high
as acrolein, acetol, formaldehyde, acetaldehyde, 1,2-propanediol temperature.98 The glycerol is evaporated and fed to the reactor
and coke. The recently published mechanisms over basic cata- diluted in an inert carrier gas, whereas the impurities remain in
lysts and over the different types of acid catalysts are the first the inert liquid. The authors describe an example of a process
attempts to explain the different product selectivities observed in which a solution of 75 wt% of glycerol with up to 3 wt%
over these types of catalysts. impurities is added to silicon-oil heated to 330 ◦ C.

This journal is © The Royal Society of Chemistry 2010 Green Chem., 2010, 12, 2079–2098 | 2095
View Article Online

To facilitate the use of crude glycerol, Dubois and Sereshiki by the use of heterogeneous catalytic systems are claimed to be
et al. instead suggest a fluidised inert solid heated to high lower than those from the conventional homogeneous process,
temperature.99,100 This leads to the evaporation of glycerol, as the leaching of the catalyst is negligible in this case. This
whereas the impurities remain inside the fluidised bed, which process is planned or already in application in six units in Europe,
can be continuously regenerated. The regeneration may consist and North & South America, with a total capacity of nearly
in a flushing step with water to dissolve inorganic salts followed 1 million tonnes per year. Nevertheless, for the near future,
by a thermal treatment to burn off the organic deposits. As both homogenous and heterogeneous processes will still have
Published on 12 November 2010. Downloaded by Pennsylvania State University on 17/09/2016 04:41:30.

an example, the authors describe a process whereby a solution to deal with the availability of different grades of crude glycerol
of 18 wt% glycerol with 2 wt% sodium chloride is put into on the global market. Therefore, the question concerning the
contact with a fluidised bed of silica particles. The fluidisation economic sense of investing (or not) in a supplementary unit for
is performed using nitrogen at 310 ◦ C. The author claims the evaporation of crude glycerol is difficult to answer at this
a separation/recovery of 99.9% of the introduced sodium stage.
chloride.
Another process enabling the use of crude glycerol was also
11 Conclusion
proposed by Dubois.101 As the evaporation of crude glycerol
at high temperature induces reactions between glycerol and The depletion of the petrol feedstock and the global warming
MONGs, this author proposed the possibility of forming acetals caused by CO2 emission have led to a change in the energy
by reacting glycerol with the acrolein in a first step (Scheme policies. Biofuels have been proposed as an alternative to fossil
14). These acetals have lower boiling points than glycerol (185– fuels for reducing the impact of greenhouse gases. Due to
240 ◦ C vs. 290 ◦ C), and thus the aforementioned undesired side- political directives, the production of bioethanol for blending
reactions are said to be reduced. The formed acetals are then with gasoline and the production of biodiesel have drastically
evaporated in the second step and subsequently dehydrated over increased. As a consequence, the market has been flooded with
an acid catalyst in the third step, to finally obtain acrolein. The crude glycerol co-produced in the biodiesel production process.
acrolein formed is partially recycled for the acetalisation in the As glycerol is a highly functionalisable molecule, many processes
first step. for its valorisation have been recently published. In particular,
the catalytic dehydration to acrolein – a precursor of sodium
polyacrylate used as a superabsorbent polymer and of DL-
methionine used for cattle feeding – is of high economic interest.
This reaction is catalysed over Brønsted and Lewis acid sites
according to two different mechanisms, which are still a matter
of debate and will undoubtedly be refined in the near future.
Furthermore, the picture of the influence of the acidity on the
catalytic performances has become clearer, showing that optimal
acidic strength is needed for high and constant performance.
As far as gas-phase dehydration is concerned, a large variety of
highly efficient catalysts has been proposed, which can be divided
into three different categories: (i) supported HPAs; (ii) zeolites;
and (iii) mixed-metal oxide type catalysts, including phosphates.
The best catalytic performance has, up to now, been achieved
with caesium salts of HPAs, yielding up to 98% acrolein in the
very early stages of the reaction. On zeolitic and mixed oxide
catalysts, the acrolein yield is generally slightly lower (around 70
Scheme 14 Process based on an intermediate acetalisation step to to 80%) and the temperature needed to reach high conversions
purify crude glycerol. is higher (around 300 ◦ C vs. 275 ◦ C for HPAs). Nevertheless, the
family of zeolites and supported HPAs offers many possibilities
In conclusion, one can see that the use of crude glycerol for tuning acidity and textural parameters, which makes possible
necessitates a supplementary upstream step at the level of fine-tuning to yield tailor-made catalysts.
the reactant evaporation. All the proposed concepts have in However, the main problem remains rapid catalyst deactiva-
common the fact that inorganic impurities like salts are retained, tion due to carbon deposition. Up to now, no catalyst has had a
whereby the poisoning of the catalyst is inhibited. Nevertheless, half-life of more than a few days, which is of course not sufficient
we have to mention that important progress has been made for commercial applications. Therefore, various approaches have
concerning the improvement of the quality of the glycerol issuing been published to regenerate the coked catalysts. These are: the
from biodiesel production processes.102 The early processes were use of a fluidised circulating catalyst, which is continuously
based on homogenous catalytic cleavage of the vegetable oils regenerated in a parallel reactor; the introduction of oxygen,
and fats in the presence of a base – for example, one of the hydrogen or sulfur dioxide into the reaction feed for continuous
new processes developed by the French company Axens uses regeneration of the catalyst inside the reactor; and discontinuous
a heterogeneous catalyst.103 This ‘Esterfip-H’ process provides regeneration by alternating the glycerol feed and the oxygen feed.
exceptionally high glycerol purities close to 98%, with only a It was even proposed to dope the catalysts with noble metals for
few traces of inorganic salts.104 Furthermore, the costs generated facilitating the splitting of the dioxygen used for regeneration. At

2096 | Green Chem., 2010, 12, 2079–2098 This journal is © The Royal Society of Chemistry 2010
View Article Online

this time, the co-feeding of oxygen seems to be the most practical 17 M. Di Serio, L. Casale, R. Tesser and E. Santacesaria, Energy Fuels,
way for extending catalyst life-time, as this technique requires 2010, DOI: 10.1021/ef901230r, in press.
18 N. O. V. Sonntag, J. Am. Oil. Chem. Soc., 1992, 75, 795.
relatively low investment and is easy compared to pulse-wise 19 Synergy Chemicals, http://www.senergychem.com.
regeneration or to a fluidised process. Nevertheless, this method 20 H. Kimura, Appl. Catal. A: Gen., 1993, 105, 147–158; H. Kimura,
is not applicable for all kinds of catalysts, as redox properties K. Tsuto, T. Wakisaka, Y. Kazumi and Y. Inaya, Appl. Catal. A:
may lead to significant decrease in the selectivity for acrolein. Gen., 1993, 96, 217–228.
21 K. Nabe, N. Izuo, S. Yamada and I. Chibata, Appl. Environ.
Liquid-phase dehydration was well studied with the rise of
Published on 12 November 2010. Downloaded by Pennsylvania State University on 17/09/2016 04:41:30.

Microbiol., 1979, 38, 1056.


research into supercritical fluids. Nevertheless, a dead-end has 22 R. Ciriminna, G. Palmisano, C. Della Pina, M. Rossi and M.
been reached, as dehydration just in supercritical water without Pagliaro, Tetrahedron Lett., 2006, 47, 6993.
23 G. W. Keulks, L. D. Krenzke and T. M. Notermann, Adv. Catal.,
acid catalyst results in low acrolein yields, while the combination 1978, 27, 183.
of supercritical conditions with acids is problematic in terms 24 M. M. Lin, Appl. Catal. A: Gen., 2001, 207, 1.
of material stress on the reaction vessel, which would lead to 25 A. Corma, G. W. Huber, L. Sauvanaud and P. O’Connor, J. Catal.,
significant investment costs. The second – and more promising 2008, 257, 163–171.
26 ICIS Market-reporter, 6 January 2010, http://www.icis.com.
– route is the use of high-boiling solvents containing solid 27 Global Industry Analysts, http://www.strategyr.com/Super_
catalysts. With this technology, a quasi-continuous process can Absorbent_Polymers_Market_Report.asp.
be achieved, even if scale-up to the industrial scale remains risky. 28 A. Yamamoto, Encyclopaedia of Chemical Technology, 3rd edn,
1978, vol. 2, p. 403.
As a consequence, the dehydration of glycerol in the liquid phase
29 M. P. Malveda, H. Janshekar, K. Yokose, CEH Marketing Research
appears to be a rather uneconomic process. Report – SRI Consulting, Major Amino Acids, June 2006.
Finally, we have shown the need for the use of crude glycerol 30 M. P. Malveda, H. Janshekar, K. Yokose, Chemical Economics
as a feedstock in regard to economic aspects. We have not Handbook-SRI Consulting, June 2006, 502.5000 B.
31 B. Katryniok, S. Paul, M. Capron and F. Dumeignil, Chem-
come across any attempt to directly use crude glycerol over SusChem., 2009, 2, 719–730.
catalytic formulations, but two promising upstream technologies 32 Schering-Kahlbaum FR695931, 1930.
have been presented, which aim to eliminate the impurities 33 H. Groll and G. Hearne (Shell), US 2042224, 1936.
of the crude glycerol during evaporation and thus avoid the 34 H. Hoyt and T. Manninen (US Ind. Chemicals. Inc.), US 2558520,
1951.
costly crude glycerol distillation. As the number of publications 35 A. Neher, T. Haas, A. Dietrich, H. Klenk and W. Girke (Degussa),
concerning this kind of technology is still very limited, it is DE 4238493, 1994.
difficult at the moment to foresee which of them is the most 36 A. Neher, T. Haas, A. Dietrich, H. Klenk and W. Girke (Degussa),
US 5387720, 1995.
promising. Furthermore, the quality of crude glycerol from 37 A. Neher and T. Haas (Degussa), US 5426249, 1995.
biodiesel production is constantly increasing, reaching a purity 38 J.-L. Dubois, C. Duquenne, W. Hoelderich and J. Kervennal
up to 98% without any inorganic salts, which makes industrial (Arkema), WO 2006087084, 2006.
applications of the proposed purification techniques rather 39 J.-L. Dubois, C. Duquenne and W. Hoelderich (Arkema),WO
2006087083, 2006.
improbable, at least in the long-term. 40 S.-H. Chai, H.-P. Wang, Y. Liang and B.-Q. Xu, J. Catal., 2007, 250,
342–349.
41 S.-H. Chai, H.-P. Wang, Y. Liang and B.-Q. Xu, Green Chem., 2007,
References 9, 1130–1136.
42 Z. H. Chen, T. Lizuka and K. Tanabe, Chem. Lett., 1984, 13, 1085.
1 United Nations Framework Convention on Climate Change, 43 T. Lizuka, K. Ogasawara and K. Tanabe, Bull. Chem. Soc. Jpn.,
http://unfccc.int. 1983, 56, 2927.
2 Directive 2003/30/EC of the European Parliament and of the 44 A. Alhanash, E. F. Kozhevnikova and I. V. Kozhevnikov, Appl.
council of 05/08/2003 on the promotion of the use of biofuels Catal. A: Gen., 2010, 378, 11–18.
or other renewable fuels for transport. 45 Y. T. Kim, K.-D. Jung and E. D. Park, Microporous Mesoporus
3 L. C. Meher, D. V. Sagar and S. N. Naik, Renew. Sustain. Energy Mater., 2010, 131, 28–36.
Rev., 2006, 10, 248–268. 46 A. K. Kinage, P. P. Upare, P. Kasinathan, Y. K. Hwang and J.-S.
4 Eurostat European Commission, http://epp.eurostat.ec.europa.eu. Chang, , Catal. Commun., 2010, 11, 620–623.
5 U.S. National Biodiesel Board, http://www.biodiesel.org. 47 S.-H. Chai, H.-P. Wang, Y. Liang and B.-Q. Xu, Green Chem., 2008,
6 Global Industry Analysts, http://www.strategyr.com/Glycerin_ 10, 1087–1093.
Market_Report.asp. 48 S.-H. Chai, H.-P. Wang, Y. Liang and B.-Q. Xu, Appl. Catal. A:
7 European Biodiesel Board, http://www.ebb-eu.org. Gen., 2009, 353, 213–222.
8 J.-L. Dubois and G. S. Patience (Arkema), WO2009044051, 2009. 49 L. Ning, Y. Ding, W. Chen, L. Gong, R. Lin, Y. Lü and Q. Xin,
9 M. J. Haas, A. J. Mc Aloon, W. C. Yee and T. A. Foglia, Biores. Chin. J. Catal., 2008, 29(3), 212–214.
Tech., 2006, 97, 671–678. 50 B. Katryniok, S. Paul, M. Capron, C. Lancelot, V. Bellière-Baca, P.
10 M. Pagliaro and M. Rossi, in The Future of Glycerol, 2nd edn (RSC Rey and F. Dumeignil, Green Chem., 2010, 12, 1922–1925.
Green Chemistry Book Series), 2010. 51 E. Tsukuda, S. Sato, R. Takahashi and T. Sodesawa, Catal.
11 C. S. Callam, S. J. Singer, T. L. Lowary and C. M. Hadad, J. Am. Commun., 2007, 8, 1349–1353.
Chem. Soc., 2001, 123, 11743. 52 H. Atia, U. Armbruster and A. Martin, J. Catal., 2008, 258, 71–
12 D. Siano, E. Santacesaria, V. Fiandra, R. Tesser, G. Di 82.
Nuzzi, M. Di Serio and M. Nastasi (ASER), WO 2006111810, 53 B.-Q. Xu, S.-H. Chai, T. Takahashi, M. Shima, S. Sato and R.
2006. Takahashi (Nippon Catalytic Chem. Ind.), WO 2007058221, 2007.
13 D. Schreck, W. Kruper, F. Varjian, M. Jones, R. Campbell, K. 54 T. Sato and R. Takahashi (Nippon Catalytic Chem. Ind.), JP
Kearns, B. Hook, J. Briggs and J. Hippler (DOW Chemicals), WO 2008088149, 2008.
2006020234, 2006. 55 H. Jo, S.-H. Chai, T. Takahashi and M. Shima (Nippon Catalytic
14 D. A. Simonetti, J. Ross-Hansen, E. L. Kunkes, R. R. Soares and J. Chem. Ind.), JP 2007137785, 2007.
A. Dumesic, Green Chem., 2007, 9, 1073. 56 S. Erfle, U. Armbruster, U. Bentrup, A. Martin and A. Brückner,
15 R. R. Soares, D. A. Simonetti and J. A. Dumesic, Angew. Chem. Appl. Catal. A, 2010, DOI: 10.1016/j.apcata.2010.04.042, in press.
Int. Ed., 2006, 45, 3982. 57 J.-L. Dubois, Y. Magatani and K. Okumura (Arkema), WO
16 BioMNC, http://www.biomnc.eu. 2009127889 and WO 2009128555, 2009.

This journal is © The Royal Society of Chemistry 2010 Green Chem., 2010, 12, 2079–2098 | 2097
View Article Online

58 X.-Z. Li (Shanghai Huayi Acrylic Acid Co), CN 101070276, 2007. 80 H. Kasuga and M. Okada (Nippon Catalytic Chem. Ind.), JP
59 A. Zhuang, C. Zhang , S. Wen, X. Zhao and T. Wu (Shanghai Huayi 2008137950, 2008.
Acrylic Acid Co), CN 101225039, 2008. 81 J.-L. Dubois (Arkema), WO 2009156664, 2009.
60 M. Okuno, E. Matsunami, T. Takahashi, H. Kasuga, M. Okada 82 Y. Arita, H. Kasuga and M. Kirishiki (Nippon Catalytic Chem.
and M. Kirishik (Nippon Catalytic Chem. Ind.), WO 2007132926, Ind.), JP 2008110298, 2008.
2007. 83 P. O’Connor, C. Corma, G. Huber and L. Savanaud (Bioecon.
61 M. Okuno, E. Matsunami, T. Takahashi and H. Kasuga (Nippon Internat. Holding), WO 2008052993, 2008.
Catalytic Chem. Ind.), JP 2007301505, 2007. 84 J.-L. Dubois (Arkema), FR 2897058, 2007.
Published on 12 November 2010. Downloaded by Pennsylvania State University on 17/09/2016 04:41:30.

62 M. Okuno, E. Matsunami, T. Takahashi and H. Kasuga (Nippon 85 H. Kasuga (Nippon Catalytic Chem. Ind.), JP 2008137952, 2008.
Catalytic Chem. Ind.), JP 2007301506, 2007. 86 S. Ramayya, A. Brittain, C. De Almeida, W. Mok and M. J. Antal,
63 C.-J. Jia, Y. Liu, W. Schmidt, A.-H. Lu and F. Schüth, J. Catal., Fuel, 1987, 66, 1364.
2010, 269, 71–79. 87 W. Bühler, E. Dinjus, H. J. Ederer, A. Kruse and C. Mas, J.
64 C.-J. Zhou, C.-J. Huang, W.-G. Zhang, H.-S. Zhai, H.-L. Wu and Supercritical Fluids, 2001, 22, 37–53.
Z. S. Chao, Stud. Surf. Sci. Catal., 2007, 165, 527–530. 88 L. Ott, M. Bicker and H. Vogel, Green Chem., 2006, 8(2), 214–220.
65 K. Pathak, K. M. Reddy, N. N. Bakhshi and A. K. Dalai, Appl. 89 V. Lehr, M. Sarlea, L. Ott and H. Vogel, Catal. Today, 2007, 121(1-
Catal. A, 2010, 372, 224–238. 2), 121–129.
66 L.-Z. Tao, S.-H. Chai, Y. Zuo, W.-T. Zheng, Y. Liang and B.-Q. Xu, 90 M. Watanabe, T. Iida, Y. Aizawa, T. M. Aida and H. Inomata,
Catal. Today, 2010, DOI: 10.1016/j.cattod.2010.03.073, in press. Biores. Technol., 2007, 98, 1285.
67 W. Suprun, M. Lutecki, T. Haber and H. Papp, J. Mol. Catal. A: 91 N. Suzuki and M. Takahashi (KAO Corp.), JP 2006290815, 2006.
Chem., 2009, 309, 71–78. 92 Y. Yoshimi, Y. Masayuki, A. Torakichi and A. Takanori (Showa
68 F. Wang, J.-L. Dubois and W. Ueda, J. Catal., 2009, 268, 260–267. Denko), JP2009179569, 2009.
69 F. Wang, J.-L Dubois and W. Ueda, Appl. Catal. A, 2010, 276, 93 A. Takanori and Y. Masayuki (Showa Denko), JP2009292773, 2009.
25–32. 94 A. Takanori and Y. Masayuki (Showa Denko), JP2009292774, 2009.
70 J.-L. Dubois (Arkema), WO2010046227, 2010. 95 E. Yoda and Y. Ootawa, Appl. Catal. A, 2009, 360, 66–70.
71 Q. Liu, Z. Zhang, Y. Du, J. Li and X. Yang, Catal. Lett., 2008, 96 M. R. Nimlos, S. J. Blanksby, X. Qian, M. E. Himmel and D. K.
127(3-4), 419–428. Johnson, J. Phys. Chem. A., 2006, 110, 6145–6156.
72 J. Deleplanque, J.-L. Dubois, J.-F Devaux and W. Ueda, Catal. 97 K. Kongpatpanich, T. Nanok, B. Boekfa and J. Limtrekul, Prepr.
Today, 2010, DOI: 10.1016/jcattod.2010.04.012, in press. Pap. Am. Chem. Soc. Div. Pet. Chem., 2010, 55(1), 115–118.
73 J.-L. Dubois (Arkema), WO2009044081, 2009. 98 J. Kijenski, A. Migdal, O. Osawaru and E. Smigiera (Inst. Chemii
74 E. Matsunami, T. Takahashi and H. Kasuga (Nippon Catalytic Przemyslowe), EP 1860090, 2007.
Chem. Ind.), JP 2007268363, 2007. 99 J.-L. Dubois (Arkema), WO 2008129208, 2008.
75 E. Matsunami, T. Takahashi and H. Kasuga (Nippon Catalytic 100 B. R. Sereshiki, S.-J. Balan, G. S. Patience and J.-L. Dubois, Ind.
Chem. Ind.), JP 2007268364, 2007. Eng. Chem. Res., 2010, 49, 1050–1056.
76 H. Redlingshoefer, C. Weckbecker, K. Huthmacher and A. Doer- 101 J.-L. Dubois (Arkema), WO2009081021, 2009.
flein (Evonik Degussa), WO 2008092533, 2008. 102 C. Perego and D. Bianchi, Chem. Eng. J., 2010,
77 H. Redlingshoefer, C. Weckbecker, K. Huthmacher and A. Doer- DOI: 10.1016/j.cej.2010.01.036, in press.
flein (Evonik Degussa), WO 2008092533, 2008. 103 R. Stern, G. Hillion, J-J. Rouxel and L. Serge (Institut Français du
78 H. Redlingshoefer, C. Weckbecker, K. Huthmacher and A. Doer- Petrol), US 5908946, 1999.
flein (Evonik Degussa), WO 2008092534, 2008. 104 L. Bournay, D. Casanave, B. Delfort, G. Hillion and J. A. Chodorge,
79 A. Ulgen and W. Hoelderich, Catal. Lett., 2009, 131, 122–128. Catal. Today, 2005, 106, 190–192.

2098 | Green Chem., 2010, 12, 2079–2098 This journal is © The Royal Society of Chemistry 2010

You might also like