Sunteți pe pagina 1din 12

DOI: 10.1002/cssc.

200900134

Towards the Sustainable Production of Acrolein by


Glycerol Dehydration
Benjamin Katryniok,[a] Sbastien Paul,[b] Mickal Capron,[a] and Franck Dumeignil*[a]
The massive increase in biodiesel production by transesterification of vegatable oils goes hand-in-hand with the availability
of a large volume of glycerol, which must be valorized. Glycerol dehydration to acrolein over acid catalysts is one of the
most promising ways of valorization, because this compound
is an important chemical intermediate used in, for example,
the DL-methionine synthesis. In this Minireview, we give a detailed critical view of the state-of-the-art of this dehydration re-

action. The processes developed in both the liquid and the gas
phases are detailed and the best catalytic results obtained so
far are reported as a benchmark for future developments. The
advances on the understanding of the reaction mechanism are
also discussed and we further focus particularly on the main
obstacles for an immediate industrial application of this technology, namely catalyst coking and crude glycerol direct-use
issues.

1. Introduction
The exhaustion of nonrenewable fossil fuels is foreseen in the
next several decades. During the last years, the production of
alternative gasolines by making use of vegetable oils and fats
has attracted the attention of many academic and industrial researchers. In addition to renewability, the resulting biodiesels
have the advantage of providing a neutral CO2 balance with a
much lower environmental impact than petrodiesels. The raw
materials for biodiesel production are vegetable oils and fats
from canola, soy, corn, or othersand a mono-alcohol (usually
methanol), which is used to cleave the fatty acids from the
glycerol backbone to finally yield the fatty-acid-esters
(Scheme 1). These acid-esters can be used as biodiesels either
directly or after blending with fossil-fuel-derived diesels.

Scheme 1. The transesterification of vegetable oils to yield the so-called biodiesels.

The use of biodiesels as substitutes or additives for petrolbased diesels can be seen as a political choice, rather than a
decision driven socio-economic reasons. Recently, this choice
has even been criticized owing to the concurrent use of the
same raw materials for food production. Nevertheless, the European Union has planned to progressively increase the biodiesel proportion in commercial fuels to reach a target of 10 %
in 2015.[1] This is a general trend, and the capacity for biodiesel
production is continuously increasing all over the world. For
example, in 2007 the USA and the EU produced 1.2 and 5.7
million tonnes (Figure 1) of biodiesel, respectively, and these
quantities are expected to double by 2012.[2]
ChemSusChem 2009, 2, 719 730

Figure 1. European biodiesel production (reproduced with permission from


the European Biodiesel Board).

This growth is accompanied by a significant increase in glycerol production, as this compound is a major byproduct of biodiesel production (ca. 10 wt %; Scheme 1). A projection of the
global production forecasts that world-wide 1.2 million tonnes
of glycerol will be generated in 2010,[3] which has to be processed to achieve a sustainable industry. Crude glycerol obtained from the biodiesel process contains 80 wt % of glycerol
but, as a drawback, also contains water, methanol, traces of
fatty acids, and various inorganic and organic compounds (re-

[a] B. Katryniok, Dr. M. Capron, Prof. F. Dumeignil


Universit des Sciences et Technologies de Lille
Unit de Catalyse et de Chimie du Solide (UMR, CNRS 8181)
Cit Scientifique, 59655 Villeneuve dAscq (France)
Fax: (+ 33) 320-436561
E-mail: franck.dumeignil@univ-lille1.fr
[b] Dr. S. Paul
Univ. Lille Nord de France, Ecole Centrale de Lille
Unit de Catalyse et de Chimie du Solide (UMR, CNRS 8181)
Cit Scientifique, BP 48, 59651 Villeneuve dAscq (France)

 2009 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

719

F. Dumeignil et al.
ferred to as MONG: matter organic non-glycerol). As a consequence, crude glycerol must in most cases be purified by an
expensive distillation step prior to further use. The proportion
of glycerol that is actually refined is now decreasing owing to
the high cost of the distillation step accompanied by a toorapid growth of the quantity of produced crude glycerol, but
mostly because of the absence of any market able to absorb
the overproduction (Figure 2). In contrast, a model based on

tion, as noted earlier the size of the existing markets is not sufficient to absorb the huge amounts of glycerol currently being
produced, and the gap between the absorption capacity of
the market and the glycerol production will steadily increase in
the near future. Today, crude glycerol that is not refined is generally burned (Figure 2), which must be considered as a dramatic waste of a potentially very interesting organic raw material. Glycerol is a molecule with a large potential for functionalization that offers many opportunities of chemical or biochemical conversions to produce value-added chemicals. A selection
of these possibilities is shown in Scheme 2 and briefly further

Figure 2. Global crude glycerol production and distillation (the difference


corresponds to the quantity of glycerol that is not upgraded and usually
burned).[7]

the production and sale of crude glycerol predicts an inverse


linear relationship between the net production costs of biodiesel and the price of glycerol.[4] New and economical ways of
using glycerol must therefore be developed to greatly increase
the demand and price of crude glycerol, thereby ensuring the
sustainability of the biodiesel sector;[5] an issue addressed by a
consortium of academic and industrial researchers recently set
up in the UK.[6]
More than 1500 direct applications of glycerol are already
known, especially in the cosmetic, pharmaceutical, and food industries.[8] The large versatility of glycerol is based on both its
chemical and physical properties. Because of its three hydroxyl
groups, glycerol is completely soluble in water and alcohols
whereas it is completely insoluble in hydrocarbons. It is a very
hydrophilic species and is employed as such when the amount
of water has to be controlled, that is, in glue or other adhesives. Furthermore, the presence of hydroxyl groups leads to
the formation of intra- and intermolecular hydrogen networks,
which explains its high boiling point (563 K at atmospheric
pressure) and high viscosity. This latter rheological property
leads to the use of glycerol as a softener in resins and plastics
but also as a lubricant, for example in pharmaceutical applications. In addition, glycerol is nontoxic and has a sweet taste. It
can be incorporated into foods, medicines, and cosmetics but,
as mentioned above, the crude glycerol from biodiesel processes contains impurities and is therefore not suitable for
these applications without a prior purification stage. In addi-

720

www.chemsuschem.org

Scheme 2. A selection of glycerol valorization pathways.

discussed below. More detailed information is available in


recent Review papers.[5, 911] The reforming of glycerol on PtRh
catalysts to yield syngas, which can be used for either the synthesis of alkanes in the FischerTropsch process or the synthesis of methanol, has been recently reported.[12, 13] A potential
application derived from this technology is the production of
hydrogen by the water-gas-shift process. Another interesting
possibility for valorizing glycerol is its selective reduction,
which leads to either propylene glycol (MPG) or to 1,3-propanediol (PD), which are used in the polymer industry. It is also
possible to halogenate glycerol to epichlorhydrin, an important
intermediate for epoxy resins, by using hydrochloric acid in the
presence of organic acids, such as caprylic acid (Solvay) or
acetic acid (Dow),[14] as catalysts working in the gaseous phase
at 353393 K under 15 bar.[15] This technology is now mature,
as Solvay has run this process commercially in France since
2007, using an existing production facility where formerly glycerol was produced from epichlorhydrin.
Another way of valorizing glycerol is etherification to glycerol-tert-butylether (GTBE), which is an excellent additive for

 2009 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

ChemSusChem 2009, 2, 719 730

Towards the Sustainable Production of Acrolein by Glycerol Dehydration


diesel blending. Furthermore, esterification of glycerol to monoacylglycerol (MAG) or diacylglycerol (DAG), used as emulsifiers
in for example foods (margarines and sauces) and cosmetics, is
also a possibility. The latter process can be catalyzed either by
a conventional alkaline catalyst or by (lipase-type) enzymes.[16]
Glycerol partial oxidation leads to a rich chemistry with many
possible products, such as glyceric, tartronic and ketomalonic
(or mesoxalic) acids, glyceraldehyde, or dihydroxyacetone
(DHA). The challenge is to find catalysts that are selective to
the target molecule among many possible products. For example, Bi/Pt catalysts were proven efficient for DHA production
with a yield of 37 % at a glycerol conversion of 70 %.[17] Oxidation can also be performed by using modified bacteria.[18] Recently, the anodic oxidation of glycerol gave a DHA yield of
25 %, which is comparable to that obtained with the aforementioned biotechnological process.[19] However, one of the most
promising ways of glycerol valorization lies in its catalytic dehydration to acrolein, which is an important intermediate for
the chemical and agricultural industries. Because of its toxicity,
acrolein is usually directly converted into derivatives such as
acrylic acid. The most important use of isolated acrolein is the
synthesis of DL-methionine via 3-methylthiopropionaldehyde
as an intermediate (Scheme 3).[20] DL-methionine is an essential

Scheme 3. Chemical pathways for the industrial manufacturing of DL-methionine.

amino acid that cannot be synthesized by living organisms. It


is widely used in meat production to accelerate animal growth.
The annual world-wide production capacity of DL-methionine
is about 500 000 tonnes.[21] Because the amounts offered by
natural methionine sources such as plants and micro-organisms are too low, it has to be industrially synthesized at an extensive scale to meet the demand. It is estimated that the
global demand will increase by 37 % in the near future, irrespective of the region.[22]
Currently, the synthesis of acrolein is based on the selective
oxidation of propene over complex multicomponent BiMoOxbased catalysts. The selectivity obtained in this process is close
to 85 % at 95 % conversion.[23] New approaches with propane
as a starting material are currently explored on a laboratory
scale, but they still suffer from insufficient yields that are incompatible with commercialization.[24] Considering the depletion issues associated with petrochemical feedstocks, sustainable resources will become more and more competitive, not to
mention the positive effect in terms of impact on the climate.
ChemSusChem 2009, 2, 719 730

In this context, obtaining acrolein from bioresources in substantial yields is an important challenge. An economical study
has shown that a competitive production of acrolein from glycerol would be possible at a glycerol price lower than ca.
US$ 300 t 1.[25] Currently (June 2009), refined glycerol still costs
US$ 400500 t 1 but crude glycerin is around US$ 100200 t 1,
which makes it a potentially very competitive raw material for
acrolein production.[26]
In this Minireview, we draw the state-of-the-art of the sustainable production of acrolein by dehydration of glycerol. Processes developed in the liquid and the gas phase are detailed,
and the best catalytic results obtained up to now are reported
and commented. The fundamentals of the reaction are also
discussed in a section devoted to its mechanism. Finally, the
main obstacles to an immediate industrial development of the
process are discussed in order to outline new insights for
future research efforts.

2. Historical Context
It has been well-known for a long time that heating glycerol
induces its decomposition into acrolein and water, with the
production of byproducts. However, an acid catalyst is needed
to have better control over the reaction and to obtain a significant yield of the unsaturated aldehyde at a moderate temperature. The first patent on this subject was published in France
in 1930.[27] The reaction was carried out in the gaseous phase
over a supported lithium phosphate catalyst, which already
provided an acrolein yield close to 75 %. Later, in 1934, Groll
and Hearne claimed a patent for the Shell company for the dehydration of an aqueous glycerol solution in the presence of
sulfuric acid at 463 K, which was supposedly the boiling point
of the solution.[28] The resulting acrolein stream was a condensed vapor with a yield of nearly 50 %. Hoyt and Manninen,
17 years later, patented the dehydration of glycerol using supported H3PO4.[29] This heterogeneous catalyst was prepared by
impregnation of a clay with 25 wt % of acid. The authors chose
petroleum oil with a boiling point higher than 573 K as a reaction medium, which enabled the use of higher reaction temperatures compared to those used by Groll and Hearne,[28] who
worked in the aqueous phase. Even when the glycerol was
used at a larger concentration (95 wt %) a high yield in acrolein
(72.3 %) could be reached. However, these early works, in both
gaseous and liquid phases, remained without further follow-up
until the end of the 20th century and the upcoming of the
massive production of cheap glycerol resulting from biodiesel
production processes.

3. Dehydration of Glycerol in the Liquid Phase


Encouraged by the increased availability of cheap glycerol,
Ramayya et al. studied the dehydration of glycerol in a solution
of sulfuric acid, and under conditions close to the critical point
of water (i.e., P = 22.1 MPa and T = 647 K).[30] (Near-)supercritical
water offers the advantage that its physical properties, such as
its dielectric constant or its ion product, can be adjusted by
varying the temperature and pressure. The mixture was heated

 2009 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

www.chemsuschem.org

721

F. Dumeignil et al.
acrolein following its formation when it is subjected to too
high temperatures. In 2007, Lehr et al. also published on the
dehydration of biomass-derived polyols in sub- and supercritical water. As far as glycerol is concerned, they reported 59 %
glycerol conversion with almost
60 % acrolein selectivity, using
Table 1. Selection of the best catalytic performances reported for the dehydration of glycerol in the liquid
ZnSO4 as a catalyst in subcritical
phase.
water conditions (see Table 1).[32]
To determine if the dehydration
T [K] P [MPa] Catalyst
Glycerol
Acrolein
Acrolein yield [%] Reactor type Ref.
of glycerol in the liquid phase
conversion [%] selectivity [%]
might offer a sustainable source
97
82
79.5
Batch
[34]
553
0.1
KHSO4
of acrolein at an industrial scale,
553

99
76
75.2
Batch
[34]

72.3
Batch
[29]
573
0.1
H3PO4/Clay
Watanabe et al. attempted to
92
81
74.0
Continuous [33]
673 34.5
H2SO4
optimize the acrolein yield.[33]
75
86
64.5
Batch
[34]
473
0.1
KHSO4
They combined the supercritical
71
87
61.8
Batch
[34]
553
0.1
KHSO4
conditions in a flow-type reactor
92
56
51.5
Batch
[34]
553
0.1
MgSO4
57
50.7
Batch
[34]
553
0.1
H3PO4 (5 wt %)/Al2O3 89
with sulfuric acid as a catalyst. It

49.0
Batch
[28]
463
0.1
H2SO4
should be noted that they con55
86
47.3
Batch
[30]
623 34.5
H2SO4
firmed the results of Bhler
50
75
37.5
Continuous [3]
633 25
ZnSO4
et al.[31] Several blank tests per40
38.0
Batch
[34]
553
0.1
H3PO4 (20 wt %)/Al2O3 95
633 34
ZnSO4
62
59
37.0
Continuous [32]
formed in the absence of a cata52
64
33.3
Batch
[34]
553
0.1
Al2ACHTUNGRE(SO4)3
lyst gave low glycerol conversions but high acrolein selectivities, with a yield up to 16 %. The
Considering the fact that sulfuric acid is soluble in supercritical
best results were obtained at 673 K under 34.5 MPa, with a sulwater, this process can be classified as homogeneous catalysis
furic acid concentration of 5 mm. In these conditions, the yield
with, as a consequence, catalyst/reaction mixture separation
of acrolein was 74 % with a glycerol conversion of around
issues. Nevertheless, the idea of using supercritical water as a
92 %. One of the main issues in this work was to sufficiently
reaction medium was a new aspect, and Bhler et al. decided
promote the acrolein formation rate to make it much larger
to deepen the subject in 2001.[31] Instead of working in a batch
than the decomposition rate, which was achieved by adding
acid in supercritical conditions.
reactor, they built a flow-type reactor and used supercritical
However, to our knowledge, the largest acrolein yield reportwater as both solvent and catalyst. They observed various deed in the liquid phase was observed at atmospheric pressure
composition products, such as acetaldehyde, formaldehyde, aland a low temperature by Suzuki et al.[34] Potassium sulfate
lylic alcohol, propionaldehyde, and, of course, acrolein. The
acrolein yield varied between 10 and 23 % depending on the
and hydrogenosulfate were used as catalysts in a batch reacreaction conditions, with the best result (23 %) obtained at a
tor, into which the glycerol solution was added dropwise to a
reaction temperature of 623 K at a pressure of 45 MPa. As
paraffin solution maintained at 553 K. The produced acrolein
aforementioned, the tests were in the absence of any catalyst
was then evaporated and recovered in an aqueous solution
or acid component, which explains the poor yields in acrolein.
containing 0.25 % of hydroquinone as stabilizer. The best yield
Recently, a combination of the ideas of Hoyt and Manniwas close to 80 %, but the described process seems difficult to
nen[29] and Ramaya et al.[30] , that is, using a supported catalyst
scale-up to an industrial level owing to its complex set-up and
and using supercritical water as a reaction medium, respectivediscontinuous processing mode.
ly, was proposed by Ott and co-workers.[3] These authors invesThus, we have seen that the dehydration of glycerol in the
tigated the dehydration of glycerol in supercritical water with
liquid phase (at atmospheric pressure or in near- or supercritizinc sulfate as a catalyst, which is surprising because this catacal conditions) leads to the formation of acrolein. However, in
lyst is supposedly not acidic enough to efficiently promote the
the absence of a catalyst, the glycerol conversion and the acroreaction. However, they justified their choice considering matelein selectivity remain very low. The reaction is efficiently catarial stress issues: water in its supercritical form is a highly corlyzed by acids, but near- or supercritical conditions induce the
rosive agent and therefore requires special and expensive steel
presence of an extremely corrosive medium, and the vessel,
grades for reactors; this corrosive power would become even
which must also resist high temperatures and pressures, must
stronger if acidic compounds were subsequently added in the
be specifically designed, resulting in high investment/maintemedium, leading to unacceptable reactor material stress. In
nance costs. The best yield claimed in these conditions is close
these conditions, the authors showed that supercritical condito 80 % for homogeneous catalysis, compared to ca. 50 % in
tions are not necessarily the optimal ones. Actually, near-critical
the best case under heterogeneous conditions. As mentioned,
water (633 K at 25 MPa) showed better conversion and selechowever, the application of homogenous catalysis on an industivity than supercritical water, that is, 75 % selectivity at 50 %
trial scale is often hampered by awkward separation issues.
conversion. This might be explained by the degradation of
Therefore, the development of a high-performance heterogein a batch reactor at 623 K, pressurized to 34.5 MPa, and acidified with 5 mm of sulfuric acid. In these conditions, and at a
residence time of 25 s, the conversion of glycerol and the selectivity towards acrolein were 55 % and 86 %, respectively.

722

www.chemsuschem.org

 2009 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

ChemSusChem 2009, 2, 719 730

Towards the Sustainable Production of Acrolein by Glycerol Dehydration


neous catalyst and/or a practical process remains a topical
issue before commercial applications of liquid-phase dehydration of glycerol to acrolein can be envisaged.

Dubois et al. published two patents on the dehydration of


glycerol over different types of catalysts.[38, 39] They carried out
several tests with zeolites, Nafion, heteropolyacids, and also
with different types of acid-impregnated metal oxides. All of
these catalysts have a well-defined HA, ranging between 9
and 18. The selectivity towards acrolein depended on the
4. Dehydration of Glycerol in the Gas Phase
catalyst. For Nafion catalysts (HA = 12) and for tungsten
4.1. Early work
oxide on zirconium oxide (HA = 14.5), the selectivity was
close to 70 % at full conversion, while the selectivity over zeoSimilar to the case of the liquid phase reaction, research on delite catalysts (HA < 2) did not exceed 60 %. The authors conhydration in the gaseous phase was abandoned until the
cluded that catalysts with a HA within the range of 10 to
1990s, when Neher et al. published two patents[35, 36] as a
16 are the best candidates.
follow-up on the old work of ScheringKahlbaum.[27] They reThe second research group that has focused its activities on
produced the tests over the lithium phosphate catalyst and
the influence of the catalysts acidity is headed by Chai. In
compared its performance to those of a set of acid catalysts
2007, they first published two papers on sustainable producwith well-defined Hammet acidities (HAs) ranging from + 3 to
tion of acrolein from glycerol. Whereas the first paper focuses
8.2. The authors claimed that aqueous solutions of 10
on niobium oxide as a catalyst,[40] the second compares the
40 wt % glycerol could be converted at 573 K over aluminasupported phosphorous acid to yield acrolein with selectivities
catalytic performance of different types of acid catalysts.[41] Beof up to 75 % at full glycerol conversion (see Table 2). The main
cause the acidity of niobium oxide strongly depends on the
byproduct was hydroxyacetone (selectivity close to 10 %). Haas
calcination temperature (the higher the calcination temperaet al. also continued the research on alumina-supported H3PO4,
ture of the hydrous niobium oxide precursor, the lower the
but with a view towards producing 1,2- and 1,3-propanediol.[37]
acidity, as previously described by Chen et al.[42] and Lizuka
et al.[43]), this compound is particularly suitable as a model cataThe yield of the intermediately formed acrolein was, however,
a bit lower (70.5 %).
lyst for studying the influence of acidity on a given reaction.
The best results were obviously obtained for catalysts calcined
at a low temperature (523573 K). In addition to their high
4.2. Catalyst acidity as an important parameter
acidity, these catalysts are also advantageous as they show
Obviously, the acidity of the active phase is a crucial parameter
higher BrunauerEmmettTeller (BET) surface areas than the
influencing the catalytic performance and stability. Its effect
catalysts calcined at higher temperatures. Nevertheless, the sehas been thoroughly studied by several groups. In 2006,
lectivity to acrolein barely exceeded 50 % at 90 % glycerol conversion. In addition, the catalysts
calcined at low temperatures
showed high carbon deposition
[a]
Table 2. Selection of catalytic performances reported for the dehydration of glycerol in the gas phase.
levels and, therefore, quick deacRef.
Active phase
Support
T [K]
Glycerol
Acrolein STY [mmol h 1 Time on
tivation. The second article of
stream [h]
conversion [%] yield [%] gcat 1]
Chais group reported the
H4SiW12O40
SiO2
548
100.0
87.0
5.3
5
[48]
screening of two dozen catalysts,
SiO2
523
100.0
85.9
1.6
5
[49]
H4SiW12O40
which they classified into four
SiO2
548
98.3
84.7
1.6
5
[47]
H4SiW12O40
different groups according to
H-MC M49

633
100.0
82.8
n.a.
n.a.
[54]
their acidity.[41] Group 1 included
H-MFI+1 % Au

633
100.0
82.8
6.3
1
[66]
ZSM-11

633
99.0
81.3
n.a.
n.a.
[54]
basic catalysts with HA values

633
100.0
80.7
6.2
3
[66]
H-MFI+0.1 % Pt
higher than 7, such as magnesib-zeolite

633
100.0
80.3
n.a.
n.a.
[54]
um oxide. These catalysts
MCM-22

633
100.0
80.1
n.a.
n.a.
[54]
showed no selectivity to acrolein
Pumice stone
673693
n.a.
80.0
n.a.
n.a.
[27]
Fe2ACHTUNGRE(PO4)3
Nd4ACHTUNGRE(P2O7)3

593
96.4
79.7
n.a.
7
[57]
at all. Group 2 contained cataMFI + Ba

633
100.0
79.0
6.2
2.5
[51]
lysts with a HA value ranging be
533
100.0
79.0
1.2
5
[60]
H2WO4
tween 3 and 7, such as zirconiZrO2/montmorillonite 533
100.0
79.0
1.2
5
[63]
WO3
um oxide. According to the work
533
100.0
77.0
1.2
5
[62]
H2WO4+1 wt % Pd
MCM-56

633
100.0
76.8
n.a.
n.a.
[54]
of Neher,[35] these catalysts
Nafion
SiO
573
100.0
76.0
19.8
7
[38]
should give high acrolein seleca-Al2O3
573
n.a.
75.0
0.3
n.a.
[35]
H3PO4
tivities but, in practice, the selecPumice stone
673693
n.a.
75.0
n.a.
n.a.
[27]
Li3PO4
tivity did not exceed 30 % al
593
98.2
74.3
n.a.
1
[57]
Gd4ACHTUNGRE(P2O7)3

593
99.6
73.9
n.a.
1
[57]
Sm4ACHTUNGRE(P2O7)3
though the performances reZSM-5

588
98.3
73.6
n.a.
n.a.
[56]
mained rather stable for 10 h
ZrO2
573
100.0
73.5
19.2
7
[38]
WO3
on-stream. Group 3, which was
ZrO2
588
100.0
70.0
12.6
4
[55]
H3PW12O40
more promising, comprised cata[a] n.a. = data not available in the publication; STY = Space-time yield of acrolein.
lysts with a HA value between
ChemSusChem 2009, 2, 719 730

 2009 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

www.chemsuschem.org

723

F. Dumeignil et al.
8 and 3. In this group, alumina-supported phosphorous
acid could be found, and also alumina-supported heteropolyacids and niobium oxide (calcined at temperatures between
673 K and 773 K). The group also contained a HZSM zeolite
and pure alumina. In good agreement with previous results,[38, 39] the selectivity was generally higher than that observed for Group 2 catalysts, with the exception of pure alumina and niobium oxide calcined at 773 K. Interesting results
were observed with the alumina-supported phosphotungstic
heteropolyacid and a mixed phase of tungsten oxide/zirconium oxide with approximately 70 % selectivity at a conversion
of 70 %, in both cases after 2 h on-stream. Unfortunately, these
catalysts showed poor stability, and their performance significantly decreased during time-on-stream. Actually, for both catalysts, the glycerol conversion dropped from 6869 % to 23
25 % between 1 and 10 h under reactant flow; the selectivity
being rather constant at roughly 6670 %. The other catalysts
were overall less efficient, and gave selectivities between 35 %
and 55 % at conversions between 55 % and 100 % after 10 h
on-stream. Group 4 contained catalysts with a HA value of less
than 8, such as Hb-zeolite, niobium oxide calcined at 623 K,
and also alumina-silicate as well as sulfonated zirconium oxide.
These catalysts were less selective to acrolein than those of
Group 3 but they showed increasing performances during
time-on-stream. For example, the glycerol conversion on an
alumina-silicate dropped from 94 to 75 % between 1 to 10 h
under reactant flow, the selectivity being rather unchanged at
roughly 4346 %. Nevertheless, similar to niobium oxide,
Group 4 catalysts were also subjected to a detrimental carbonization, with a carbon deposit that represented between
100 mg and 400 mg of carbon per gram of catalyst after 10 h
on-stream.
The same team further published two other papers in 2008
and 2009 dealing with zirconia- and silica-supported 12-tungstophosphoric acid catalysts.[44, 45] These studies showed that
the nature of the catalyst support has a significant effect on
the thermal stability and on the dispersion of the Keggin-type
active phase. The use of ZrO2 as a support led to better results
as far as the deactivation of the catalyst is concerned. The
Keggin-anion density at the surface of the support was identified as a key parameter for tuning the activity and the selectivity of the HPA for acrolein production, which is in good agreement with the conclusions of Ning et al. who used activated
carbon-supported silicotungstic acid catalysts.[46] The latter
claimed that a 10 wt % H4SiW12O40 supported catalyst gives the
best acrolein space-time yield ever reported in the literature
(i.e., 68.5  10 3 molacrolein h 1 gactive phase 1). This promising performance was attributed to the good dispersion of the HPA on
the support surface and also to the relative quantities of
strong acid sites.
In this section we have seen that the tuning of the acidity of
the active phase is of prime importance for getting a good catalytic performance. However, this characteristic alone is not
sufficient to also get a sufficient stability of the active phase
and a good selectivity to acrolein. It seems that the properties
of the support and the way the active phase is dispersed on

724

www.chemsuschem.org

can also have large consequences. This point is discussed in


section.
4.3. Influence of the textural properties: A strong effect
This very interesting aspect has been recently discussed by
Tsukuda and co-workers.[47] This group works with silica-supported heteropolyacids and focuses on the influence of the
textural properties of the porous support on the selectivity.
They used three silicas with pore sizes of 3, 6, and 10 nm. The
active phase was either a heteropolyacid or a conventional inorganic acid, such as phosphoros or boric acid. Good results
were obtained when using Keggin-type heteropolyacids such
as phosphotungstic acid (H3PW12O40) or silicotungstic acid
(H4SiW12O40) on the 6 nm silica support. At full conversion, the
selectivities towards acrolein were 65 % and 75 %, respectively,
whereas the molybdenum-based homologous heteropolyacid
(i.e., H3PMo12O40) did not yield a high selectivity (34 %). These
values could be increased even further by changing the support to a silica with a larger pore diameter. The maximum selectivity measured was then 86 % for the silicotungstic heteropolyacid deposited on a 10 nm pore diameter silica with a
30 wt % loading (at full glycerol conversion). For very small
pore diameters (i.e., 3 nm), the selectivity remained nearly unchanged with time-on-stream whereas the conversion decreased by a factor of two, probably because of coking related
to steric limitations. The influence of temperature was also
studied and the optimum, in terms of acrolein yield, was
548 K.
These results show that the size of the pores has a direct influence on the stability of the catalyst. Actually, if steric limitations do not enable the rapid desorption and diffusion of the
products in the porous network of the catalyst, then coking is
more likely to occur, which results in blocking of the access to
the active phase and, thus, leads to the deactivation of the catalyst.
4.3. The combined effect of acidity and porous structure
The above-mentioned studies point out two important parameters that govern the catalytic performances: the acidity and
the textural properties of the catalyst. These two parameters
were concomitantly investigated by Chai and Tsukuda, who
gathered their efforts, resulting in the publication of two
common patents.[48, 49] These patents actually focus on supported heteropolyacids as catalysts and on their catalytic performances in relation to the textural properties of the porous support. They carried out a large number of tests, varying the
active catalyst phase, loading amounts, and supports. As previously reported, phosphomolybdic acid was unselective compared to phosphotungstic acid and silicotungstic acid.[47] They
obtained the largestto our knowledgeacrolein selectivity
ever reported in the literature using a catalyst comprising 20
30 wt % silicotungstic acid deposited onto porous silica with a
pore diameter of 10 nm (i.e., 87 % at 548 K). Smaller pore diameters (3 nm) or higher reaction temperatures (573 K) always led
to a decrease of the catalytic performance indicators (selectivi-

 2009 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

ChemSusChem 2009, 2, 719 730

Towards the Sustainable Production of Acrolein by Glycerol Dehydration


ty, conversion) by a factor two. The influence of the BET specific surface area (SSA) of the support was also studied but
proved to have little effect. They compared two silicas with
identical pore diameter but with different SSAs. A twofold increase of the surface area did not result in any significant
effect on the catalytic performance.
Apparently, the use of heteropolycompounds as active
phase offers an easy way for controlling the acidity, and thereby the catalytic performance, by carefully tuning the composition of the Keggin HPA. Encouraged by the work of Tsukuda
et al. on heteropolyacids,[47] Atia et al. published a comparison
of the performances of silicotungstic acid and phosphomolybdic acid supported on alumina, silica, or aluminosilicates.[50]
The samples, prepared by incipient wetness impregnation,
were calcined at 673 K for 4 h prior to use and their acidity
was determined by ammonia temperature-programmed desorption. The silica-supported heteropolycompound was more
acidic than the alumina-supported one. Moreover, the acidity
of the alumina-supported species was similar to that of pure
alumina, which was explained by the substitution of alumina
Brnsted sites by HPA ones. The reaction conditions for the
catalytic tests were chosen similar to those used by Tsukuda,[47]
for the sake of easier direct comparison. Irrespective of the
type of alumina-supported catalyst, total glycerol conversion
was observed for a temperature higher than 548 K but significant differences in selectivity were observed (i.e., between 30
and 60 %), underlining the strong influence of the support. The
best results were obtained for silicotungstic acid supported on
alumina and aluminosilicates with acrolein yields of 63 % and
67 %, respectively, but the authors preferred to rule out any
conclusion about silica-supported silicotungstic acid because
the deactivation was too fast and reliable data could not thus
be acquired. They did indicate the instantaneous initial selectivity over silicotungstic acid supported on a silica with a mean
pore diameter of 11 nm, which was 73 %. The contact time was
further varied by flow adjustment, and full conversion was achieved for values larger than 0.2 s, which illustrates the high reactivity of glycerol. Although acrolein is subjected to degradation by consecutive reactions, the selectivity did not drastically
decrease with the increase in contact time from 0.15 to 0.7 s.
4.5. Recent progress: Other classes of active catalysts
The use of supported heteropolycompounds is not the only
way to reach good catalytic performance, as demonstrated by
Dubois et al. in recent work.[38] Actually, zeolitic structures are
also promising and have been well-studied for three years.
In 2007, Okuno and coworkers patented metallosilicate catalysts with an MFI structure; a typical structure of zeolites with
a 3D pore network.[51] These catalysts are prepared in a fourstep synthesis, starting from the impregnation of a silica support with a solution containing a metal component such as
aluminum hydroxide, tetraalkylammonium components such
as tetramethyl- or tetraethylammonium, and a certain amount
of an alkaline compound. This leads to a precursor with the
formula Si1TxLyACHTUNGRE(SDA)z, where T is the metal atom, L is the alkaline, and SDA the tetraalkyl component. In the second step,
ChemSusChem 2009, 2, 719 730

this precursor is crystallized by hydrothermal treatment to give


a metallosilicate of MFI structure. In the third step, the introduced alkali atoms are substituted by ammonia in an ion-exchange reaction before the final catalyst is obtained in the
fourth step, which is called the burning step and means calcination at a temperature of up to 873 K.
The catalytic performance tests revealed good results for
aluminosilicates and gallosilicates whereas ferrosilicates were
less selective. The aluminosilicates offered even stable performance without addition of any alkali atom or ion-exchange,
with a selectivity for acrolein of around 65 %. This selectivity
did not change significantly with the aforementioned addition/
exchange, which was also observed over the gallosilicates. In
this patent, the results claimed over zeolitic catalysts are quite
similar to those reported in the work by Dubois et al. (i.e., 70 %
acrolein selectivity at glycerol full conversion).[38] In further
publications, the same group reported tuning of the acidity by
modifying the Si/Al ratio of the zeolite, and the best results
were achieved using a ratio of 28 with a 71.2 % yield of acrolein.[52, 53]
A second group, headed by Li, specifically studied aluminosilicates catalysts.[54] The authors reported about protonated zeolites, for example, MCM-49, which was synthesized under hydrothermal conditions and further protonated by treatment
with ammonium nitrate before calcination. This procedure
seems very similar to those reported by Okuno et al.[51] and
Dubois et al.,[38] but the catalytic performances observed by Li
et al. are better. The selectivity to acrolein reached a value
close to 83 % at full glycerol conversion, which is the best
result ever published for a zeolitic catalyst (see Table 2). This
might be due to the optimization of the reaction parameters
such as temperature and contact time.
Zhou and co-workers proposed a synthesis procedure using
dual templates of micro- and mesoporous ZSM-5 composites
used as catalysts in the dehydration of glycerol to acrolein.[56]
The best result achieved with these catalysts was 73.6 % acrolein selectivity at 98.3 % glycerol conversion, which is a good
performance. Rare-earth pyrophosphates were also claimed by
Liu and co-workers to be efficient catalysts for the production
of acrolein by vapor-phase dehydration of glycerol[57] with an
acrolein yield as high as 80 % achieved over Nd4ACHTUNGRE(P2O7)3. Matsunami et al. also reported results on silica-supported phosphates
doped by alkali and metal salts but in this case, the acrolein
yield did not exceed 67 %.[58, 59] More recently, Redlingshofer
et al. proposed tungstates as efficient catalysts for this reaction.[60, 61] The reaction was carried out at a relatively low temperature (533 K) and exhibited good performances with an
acrolein yield between 77 and 79 %. Nevertheless, the catalysts
tended to deactivate under the stream, whereby the acrolein
yield decreased at a rate of 5 % per 10 h. Additionally, the authors described the possibility to regenerate the catalyst by
oxygen treatment at 623 K for 5 h and claimed to recover the
initial catalytic performance after this treatment.[62, 63]
As discussed above and further summarized in Table 2, a
great variety of active phases can be used to selectively dehydrate glycerol to acrolein in the gas phase. One key parameter
is the tuning of the solid acidity, but the physical properties of

 2009 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

www.chemsuschem.org

725

F. Dumeignil et al.
the catalyst are also of prime importance to achieve good performances. Here, the supports can play a major role through
modifying the textural properties of the catalyst, which has a
dramatic effect on the selectivity probably because of steric
and diffusion issues. The best results published so far were obtained at a relatively low temperature (523548 K) using silicasupported HPAs. These catalysts, however, tend to deactivate
quite rapidly with time-on-stream. The zeolite family is also
promising but needs higher working temperatures (more than
623 K) to achieve high yields.
When the necessary data were available in the papers, we
calculated the acrolein space-time yield (STY) taking the total
catalyst mass as a reference. The obtained values are reported
in Table 2. Accordingly, the best result obtained up to now
were achieved by Dubois et al. on a SiO2-supported Nafion catalyst.[38] The STY was 20  10 3 molacrolein h 1 gcatalyst 1. However,
another method to calculate the STY takes into account only
the active phase mass, and when using this method the best
result was achieved by Ning et al. with an activated carbonsupported silicotungstic acid catalyst (10 wt % H4SiW12O40).[64]
The STY was 68.5  10 3 molacrolein h 1 gactive phase 1.

cation, the only possible product after cleavage of H3O + is


acrolein, whereas either acrolein or acetaldehyde and formaldehyde can be formed starting from the primary carbocation.
The radical pathway (Scheme 4 b) starts with the abstraction of
a hydrogen from a primary carbon by an OHC radical. The resulting radical species further loses an OHC, which leads to the
formation of acrolein.
Tsukuda et al. subsequently proposed a more formal reaction framework (Scheme 5).[47] The first step, glycerol dehydration, leads to the formation of two enols, which are in tauto-

5. Reaction Mechanism
Whereas the research for an efficient catalyst can follow a
more or less purely applied approach, an understanding of the
reaction mechanism is a more fundamental aspect. This implies
the identification of the intermediate steps and the explanation of formation of the byproducts. A first proposal for the
glycerol activation in near- or supercritical conditions was
made by Bhler et al.[31] Two pathways were claimed, via either
an ionic or a radical mechanism. The ionic reaction
(Scheme 4 a) begins with the protonation of glycerol on either
one primary hydroxyl group or the secondary hydroxyl group.
Afterwards, the elimination of a water molecule leads to the
formation of a carbocation. In the case of the secondary carbo-

Scheme 4. Reaction pathways to the formation of acrolein considering (a) an


ionic and (b) a radical pathway.[31]

726

www.chemsuschem.org

Scheme 5. Mechanism proposed by Tsukuda et al. and by Chai et al.[41, 47]

meric equilibrium with the corresponding ketone (hydroxyacetone) and aldehyde (3-hydroxypropionaldehyde). The latter
subsequently reacts in a second step either via dehydration to
acrolein or via a retroaldol reaction to give formaldehyde and
acetaldehyde, which may be easily oxidized to acetic acid (in
the presence of O2). As an important conclusion, Tsukuda, in
agreement with Bhlers previous proposal,[31] pointed out that
the key to obtaining a high selectivity for acrolein lies in the
control of the first dehydration step to favor the formation of
3-hydroxypropionaldehyde while suppressing the formation of
hydroxyacetone, which is identified as the main byproduct of
the process.
The same mechanism has been further completed by Chai
et al. (Scheme 5),[41] who added the two hydrogenation steps
that can occur starting from acrolein and hydroxyacetone to
form allylic alcohol and 1,2-propanediol, respectively, which are
sometimes actually observed. The hydrogenation of the formaldehyde to methanol or its decomposition into CO+H2 is also
mentioned.
These schemes are interesting for a first approach, but an
even deeper analysis is needed to explain the formation of all
the observed byproducts. Corma postulated a more complex

 2009 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

ChemSusChem 2009, 2, 719 730

Towards the Sustainable Production of Acrolein by Glycerol Dehydration


reaction framework (Scheme 6).[25] The first dehydration step is
obviously the same as that previously described by Tsukuda
and Chai, and leads to hydroxyacetone or 3-hydroxypropional-

were found, which is explained by the formation of cyclic C6compounds identified by GC-MS analysis (Scheme 7 a). The authors further identified various furan derivates (Scheme 7 b),
which must be thus taken into account when one specifically
seeks for fine information on the minor reaction byproducts.

Scheme 6. Reaction network proposed by Corma et al.[25]

dehyde, which further reacts in a second dehydration step to


yield acrolein.[47] To go further and investigate the secondary
reactions, the authors directly used hydroxyacetone as a reactant at 623 K and classified the main reaction products in three
groups: acids (9.4 %), other aldehydes (52.0 %), and coke
(27.7 %). They also observed the formation of low amounts of
acetone (4.0 %), acetaldehyde (1.4 %), and carbon monoxide
(1.2 %). Nevertheless, the conversion of hydroxyacetone did
not exceed 25 %, which means that this compound is relatively
stable and is thus usually identified as the major byproduct of
the reaction of dehydration of glycerol to acrolein.
To investigate the consecutive reactions of acetone, they repeated this method by reacting acetone at 623 K. The resulting
product distribution is dominated by unsaturated hydrocarbons such as butene (27.9 %), propylene (1.7 %), C5+C68 aromatics (6.6 %), and coke (31.9 %; note that the latter is the
known consecutive product of unsaturated hydrocarbons). The
other identified products were various acids (20.2 %) and other
aldehydes (7.3 %) at an overall conversion that did not exceed
14 %. With their findings, they thus completed the reaction
scheme proposed by Chai and Tsukuda by adding further sidereactions originating from hydroxyacetone. Thereby, they explained the formation of oligomers and coke by consecutive
reactions starting from hydroxyacetone via acetone and acetaldehyde.
Suprun et al. chose a similar approach to investigate the
consecutive reactions of 3-hydroxypropionaldehyde and hydroxyacetone.[65] They could confirm the retroaldol reaction
previously postulated by Tsukuda,[47] which leads to the formation of formaldehyde and acetaldehyde issued from 3-hydroxypropionaldehyde. In addition, significant amounts of coke
ChemSusChem 2009, 2, 719 730

Scheme 7. Products derived from (a) 3-hydroxypropionaldehyde and (b) hydroaxyacetone.[65]

6. Catalyst Deactivation: A Key Issue


It has been reported above that very efficient catalysts for the
dehydration of glycerol into acrolein can be prepared. Unfortunately, these catalysts are also very unstable under reaction
conditions and prove to deactivate very quickly, most probably
because of the formation of coke, which makes their straightforward use in an industrial plant difficult at this time. Therefore, solutions have to be found to avoid or limit the deposition of coke onto the catalyst surface or at least to optimize
the regeneration of the catalysts.
At least three kinds of solutions have been proposed to continuously regenerate the catalyst: (1) co-injection of oxygen
with the gas feed to yield in situ regeneration (co-feeding),[39]
(2) cyclic regeneration of the used catalyst by injection of a
flow/pulses of air or oxygen,[67] and (3) circulation of the catalyst in a moving bed reactor with regeneration in a parallel
vessel (as for the FCC process).[25] Whereas the first option is associated with the risk of generating explosive conditions and/

 2009 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

www.chemsuschem.org

727

F. Dumeignil et al.
or oxidizing the various reaction products, the second alternative is accompanied by the disadvantage of a loss of productivity. The third option does not suffer from these drawbacks but
the construction and the operation of a circulating bed reactor
implies serious technological difficulties that can only be overtaken by skilled personnel and fine process control.
The idea of working with air in the reaction gas was first
proposed by Dubois.[38] To stay out of the explosivity range,
the oxygen fraction must never exceed 7 vol %. Therefore, the
composition of the reaction feed was adjusted to 6 % of
oxygen for 4.5 % of glycerol, the remaining 89.5 % being
steam. Thereby, the author claimed to reduce catalyst deactivation and even to inhibit the formation of byproducts such as
hydroxyacetone in these conditions. Kasuga et al. injected air
into the gas feed to have a feed ratio of 7 % of oxygen for
27 % of glycerol, the remaining 66 % being steam and nitrogen
from air. Nevertheless, the selectivity towards acrolein was
then rather low and did not exceed 45 % after 24 h under reactant flow.[67] To optimize the regeneration, Kasuga et al. also
proposed to modify their MFI protonated-zeolite with a small
amount of metal (Pt, Pd, Ru, Cu, Ir, or Au). By this doping, they
claimed to accelerate the split of the dioxygen used for the regeneration.[66] The best results were obtained with 0.1 wt % Pt
and 1 wt % Au (80.7 and 79.7 % of acrolein yield, respectively,
at full glycerol conversion after 150 min under reaction conditions).
The possibility of periodic regeneration was studied by Arita
et al., who regenerated a used H-ZSM5 catalyst under air flow
and showed the possibility of recovering the initial performance. However, the hot-spot temperature during regeneration surpassed the regeneration temperature by more than
100 K.[67] Atia et al. carried out long-term runs (up to 300 h)
and observed that the selectivity to acrolein remained stable
whereas glycerol conversion decreased more or less linearly.
This effect could be well explained by the deposit of coke
onto the catalyst surface and the concomitant decrease of the
number of accessible active sites. To verify this hypothesis,
coked catalysts were regenerated under a flow of 1 % oxygen
in nitrogen at 598 K during 24 h. After this treatment, the catalyst performed identical to the fresh catalyst. To quantify this
effect, the used catalyst was also analyzed by thermogravimetry, which confirmed the carbon deposit hypothesis.[50]
Meanwhile, Corma et al. adopted the idea of a circulating
bed reactor.[25, 68] The authors studied the opportunity of injecting crude glycerol directly into FCC plants. An advantage
would be the use of existing facilities and, therefore, no need
for investments for building specific infrastructures. It is worth
noting that this idea has already been proposed in a different
form by Dubois, who pointed out the opportunity of injecting
glycerol into propylene oxidation plants to concomitantly yield
acrolein from both sources.[69] In the FCC plants, the heat recovered by the burning of coke could be used to provide the
energy necessary for the evaporation of glycerol. Corma et al.
concluded that an autothermal process is possible in that case.
Nevertheless, the authors did not limit their view on the dehydration but also investigated the possibilities for reforming the
glycerol at 773 K to 873 K to produce ethylene and propylene.

728

www.chemsuschem.org

Thereby, the glycerol feedstock may be an alterative or a complement to naphta cracking.


An original approach was followed by Kasuga et al., who
used some pretreatment of the catalysts (protonated MFI zeolite or Alox alumina).[70] Three types of pretreatments were
tested: (1) under a flow of hydroxyacetone, water, and nitrogen; (2) under a flow of acrolein, water, and nitrogen; and
(3) under a flow of hydroxyacetone, water, and air. Irrespective
of the pretreatment method, the Alox catalyst always led to
very poor acrolein yields while over the MFI catalyst the pretreatments enabled higher selectivities in the first minutes
under the reactant flow. However, after 150 min on-stream, the
catalytic performances recovered values similar to those obtained without preliminary pretreatment, which means that
this effect is only effective in the early stage of the reaction.
Nevertheless, in this work the authors underlined the possibility of recovering the initial performance by regenerating the
catalysts at 773 K under air flow.

7. Direct Use of Crude Glycerol as Reactant:


An Important Economic Issue
The price of glycerol drastically depends on the required technical grade. Whereas refined glycerol currently costs between
US$ 500 and 600 t 1 (June 2009), crude glycerol can be obtained for only US$ 100200 t 1.[26] Nevertheless, the glycerol
market has been highly volatile in 20072008, with glycerol
costs of up to US$ 1800 t 1 for refined glycerol and up to
US$ 500 t 1 for crude glycerol.[26] It should be mentioned that
the availability of glyceroland therefore its pricealso depends on political decisions concerning biodiesel production
and agricultural subsidies. Therefore, enabling the use of crude
glycerol is a crucial issue for achieving a sustainable and economically viable production of acrolein. However, as mentioned earlier crude glycerol is generally contaminated with byproducts issued from biodiesel process (inorganic traces such
as water, sodium and potassium carbonate, and organic traces
such as esters, fatty acids, and alcohols). The use of crude glycerol as a feedstock may therefore cause problems either by
poisoning the catalyst or by causing plugs because of the deposition of high-boiling-point organic materials or inorganic
salts. From our knowledge, at this date no catalytic test using
this type of feedstock has been reported in the literature. All
the disclosed works were carried out using diluted aqueous
solutions prepared from refined glycerol. Furthermore, it has
been found that the acrolein yield strongly depends on the
glycerol concentration in the reactor feed, with a decrease of
the yield with increasing concentration.[35, 36] However, the use
of crude glycerol is of prime economical importance and different concepts have already been proposed to avoid the costly
distillation of crude glycerol and to overtake the obstacles
raised by its use as a feedstock. Kijenski et al. proposed a
modified evaporation system in which the crude glycerol is
brought into contact with an inert liquid at high temperature.[71] The glycerol is evaporated and fed into the reactor by
an inert carrier gas while the impurities remain in the inert
liquid. The author describes a process example where a solu-

 2009 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

ChemSusChem 2009, 2, 719 730

Towards the Sustainable Production of Acrolein by Glycerol Dehydration


tion of 75 wt % of glycerol with up to 3 wt % impurities was
added to a silicon-oil heated at 603 K. The glycerol vapors
were then carried by a nitrogen flow to feed the reactor.
To facilitate the use of crude glycerol, Dubois rather suggested a fluidized inert solid heated at high temperature.[72] This
leads to the evaporation of glycerol vapors while the impurities remain inside the fluidized bed, which can be continuously
regenerated. The regeneration may consist in a flushing step
with water to dissolve inorganic salts followed by a thermal
treatment to burn organic deposits. The author describes an
example process in which a solution of 18 wt % of glycerol
with 2 wt % of sodium chloride is fed into a fluidized bed of
silica particles. The fluidization is performed using nitrogen at
583 K. As a result, the author claimed a recovery of 99.9 % of
the introduced sodium chloride.

8. Conclusions
The finite nature of the fossil fuel supply and the global warming problem have led to a change in energy policies. In this
context, biofuels have received increased attention as CO2-neutral alternatives to crude oils. Beside bioethanol as a substitute
for gasoline, the use of biodiesel has drastically progressed
especially in the European Union. The increase in biodiesel
production has also led to an oversupply of glycerol as a coproduct. Because glycerol is a highly functionalizable molecule,
many processes for its valorization have been recently published. In particular the catalytic dehydration to acroleina
precursor of DL-methionine used for cattle feedingis of economic interest.
This reaction can proceed in either the gas or the liquid
phase. The reaction conditions in the liquid phase include the
possibility to use supercritical solvents. In terms of economical
interest, the use of heterogeneous catalysis is usually preferred
rather than the use of homogenous catalysis, as the separation
of the catalyst from the reaction medium is simplified. Therefore, the possibility of using supported acids instead of dissolved ones is preferable, even if it has been demonstrated
that both technologies show comparable catalytic performances. A second important step on the road to commercialization
is demonstrating the feasibility of working in continuous
mode. Experimental results have already been published, and
show only slightly decreased performances when compared
with discontinuous experiments.[33] At last, the use of supercritical water as a reaction medium has been proven to be adaptable to the reaction. Nevertheless, the combination of supercritical and acid conditions leads to high demands on the materials used for the reaction vessel and all peripherals. For all
these reasons, and even if an acrolein yield of 80 % has already
been achieved in the liquid phase, the possibility of a commercial application seems questionable at this date.
Concerning the gas-phase reaction, this Minireview has outlined a great variety of efficient catalysts. We can classify them
according to three different categories: (1) supported inorganic
acids, with a subcategory for supported heteropolyacids;
(2) protonated zeolites; and (3) mixed-oxide type catalysts. The
best acrolein yields are, up to now, achieved with supported
ChemSusChem 2009, 2, 719 730

inorganic acids and are in the range of 75 to 87 %. On zeolitic


and mixed-oxide catalysts, the acrolein yield is generally slightly lower (around 70 to 80 %) and the temperature needed to
reach high conversions is higher (around 573623 K against
523573 K for the inorganic acids). For supported heteropolyacids, the published results already give a clear image of the
properties required to design efficient catalysts. These are a
controlled proper acidity of the active phase, and appropriate
textural properties (mainly porosity) brought by the support to
avoid steric and diffusion limitations effects. Applying these
two concepts, an acrolein yield of 87 % has already been achieved.
Nevertheless, the main problem remains the rapid catalyst
deactivation owing to carbon deposition. Up to now, no catalyst exceeds a half-life time of a few days, which is necessary
for commercial applications. Different approaches have been
published to regenerate the coked catalyst. These include the
use of a fluidized circulating catalyst that is continuously regenerated in a parallel reactor, the introduction of air or
oxygen into the reaction feed for a continuous regeneration of
the catalyst directly inside the reactor, and the discontinuous
regeneration by alternating the glycerol feed and the oxygen
feed. Modifying the catalyst formulations to facilitate the splitting of the dioxygen used for regeneration has also been proposed. Currently, none of these three has proven technologically superior to the others.
Finally, we have shown the stringent need for using crude
glycerol as a feedstock with regard to economical aspects. We
did not find any attempt of directly using crude glycerol over
catalytic formulations but two promising processes have been
presented, which intend to eliminate the impurities of the
crude glycerol during evaporation and avoid costly crude glycerol distillation. As in the case of the catalyst regeneration,
none of these processes have proven superior to the others at
this moment.

Acknowledgements
The authors would like to thank the French Research Network
No. 1 (RDR1) of the CPDD program as well as CE (through contract number MIRG-CT-2007-046 383) for their financial support
for realizing this bibliographic study.
Keywords: acrolein biomass catalysis dehydration
glycerol
[1] Directive 2003/30/EC of the European Parliament and of the council of
05/08/2003 on the promotion of the use of biofuels or other renewable
fuels for transport.
[2] Eurostat European Commission, http://epp.eurostat.ec.europa.eusion.
[3] L. Ott, M. Bicker, H. Vogel, Green Chem. 2006, 8, 214.
[4] M. J. Haas, A. J. Mc Aloon, W. C. Yee, T. A. Foglia, Bioresour. Technol.
2006, 97, 671.
[5] M. Pagliaro, M. Rossi, The Future of Glycerol: New Uses of a Versatile Raw
Material, RSC Green Chemistry Book Series, 2008.
[6] http://theglycerolchallenge.org/index.htm.
[7] Data sources: European Biodiesel Board and HB I.

 2009 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

www.chemsuschem.org

729

F. Dumeignil et al.
[8] C. S. Callam, S. J. Singer, T. L. Lowary, C. M. Hadad. , J. Am. Chem. Soc.
2001, 123, 11743.
[9] G. Paulo da Silva, M. Mack, J. Contiero, Biotechnol. Adv. 2009, 27, 30.
[10] Y. Zheng, X. Chen and Y. Shen, Chem. Rev. 2008, 108, 5253.
[11] M. Pagliaro, R. Ciriminna, H. Kimura, M. Rossi, C. Della Pina, Angew.
Chem. 2007, 119, 4516; Angew. Chem. Int. Ed. 2007, 46, 4434.
[12] D. A. Simonetti, J. Ross-Hansen, E. L. Kunkes, R. R. Soares, J. A. Dumesic,
Green Chem. 2007, 9, 1073.
[13] R. R. Soares, D. A. Simonetti, J. A. Dumesic, Angew. Chem. 2006, 118,
4086; Angew. Chem. Int. Ed. 2006, 45, 3982.
[14] D. Siano, E. Santacesaria, V. Fiandra, R. Tesser, G. Di Nuzzi, M. Di Serio, M.
Nastasi (ASER), WO2006111 810, 2006.
[15] D. Schreck, W. Kruper, F. Varjian, M. Jones, R. Campbell, K. Kearns, B.
Hook, J. Briggs, J. Hippler (DOW Chemicals), WO2006020 234, 2006.
[16] N. O. V. Sonntag, J. Am. Ceram. Soc. 1992, 75, 795.
[17] H. Kimura, Appl. Catal. A: Gen. 1993, 105, 147; H. Kimura, K. Tsuto, T. Wakisaka, Y. Kazumi, Y. Inaya, Appl. Catal. A 1993, 96, 217.
[18] K. Nabe, N. Izuo, S. Yamada, I. Chibata, Appl. Environ. Microbiol. 1979,
38, 1056.
[19] R. Ciriminna, G. Palmisano, C. Della Pina, M. Rossi, M. Pagliaro, Tetrahedron Lett. 2006, 47, 6993.
[20] A. Yamamoto, Encyclopedia of Chemical Technology 3rd ed., 1978, Vol. 2,
403.
[21] M. P. Malveda, H. Janshekar, K. Yokose, CEH Marketing Research Report
SRI Consulting, Major Amino Acids, June 2006.
[22] M. P. Malveda, H. Janshekar, K. Yokose, Chemical Economics HandbookSRI Consulting, June 2006, 502.5000 B.
[23] G. W. Keulks, L. D. Krenzke, T. M. Notermann, Adv. Catal. 1978, 27, 183.
[24] M. M. Lin, Appl. Catal. A: Gen. 2001, 207, 1.
[25] A. Corma, G. W. Huber, L. Sauvanaud, P. OConnor, J. Catal. 2008, 257,
163.
[26] Petrosil Glycerine Report, June 12, 2009, www.glycerinereport.com.
[27] Schering-Kahlbaum AG, FR 695931, 1930.
[28] H. Groll, G. Hearne (Shell), US 2042224, 1936.
[29] H. Hoyt, T. Manninen (US Ind. Chemicals. Inc.), US 2558520, 1951.
[30] S. Ramayya, A. Brittain, C. DeAlmeida, W. Mok, M. J. Antal, Fuel 1987, 66,
1364.
[31] W. Bhler, E. Dinjus, H. J. Ederer, A. Kruse, C. Mas, J. Supercrit. Fluids
2002, 22, 37.
[32] V. Lehr, M. Sarlea, L. Ott, H. Vogel, Catal. Today 2007, 121, 121.
[33] M. Watanabe, T. Iida, Y. Aizawa, T. M. Aida, H. Inomata, Bioresour. Technol.
2007, 98, 1285.
[34] N. Suzuki, M. Takahashi (KAO Corp.), JP 2006290815, 2006.
[35] A. Neher, T. Haas, A. Dietrich, H. Klenk, W. Girke (Degussa), DE 4238493,
1994.
[36] A. Neher, T. Haas, A. Dietrich, H. Klenk, W. Girke (Degussa), US 5387720,
1995.
[37] A. Neher, T. Haas (Degussa), US 5426249, 1995.
[38] J.-L. Dubois, C. Duquenne, W. Hoelderich, J. Kervennal (Arkema), WO
2006087084, 2006.
[39] J.-L. Dubois, C. Duquenne (Arkema), WO 2006087083, 2006.
[40] S.-H. Chai, H.-P. Wang, Y. Liang, B.-Q. Xu, J. Catal. 2007, 250, 342.
[41] S.-H. Chai, H.-P. Wang, Y. Liang, B.-Q. Xu, Green Chem. 2007, 9, 1130.
[42] Z. H. Chen, T. Lizuka, K. Tanabe, Chem. Lett. 1984, 13, 1085.

730

www.chemsuschem.org

[43]
[44]
[45]
[46]
[47]
[48]
[49]
[50]
[51]
[52]
[53]
[54]
[55]
[56]
[57]
[58]
[59]
[60]
[61]
[62]
[63]
[64]
[65]
[66]
[67]
[68]
[69]
[70]
[71]
[72]

T. Lizuka, K. Ogasawara, K. Tanabe, Bull. Chem. Soc. Jpn. 1983, 56, 2927.
S-H. Chai, H-P. Wang, Y. Liang, B-Q. Xu, Green Chem. 2008, 10, 1087.
S.-H. Chai, H.-P. Wang, Y. Liang, B.-Q. Xu, Appl. Catal. A 2009, 353, 213.
L. Ning, Y. Ding, W. Chen, L. Gong, R. Lin, Y. L, Q. Xin, Chin. J. Catal.
2008, 29, 212.
E. Tsukuda, S. Sato, R. Takahashi, T. Sodesawa, Catal. Commun. 2007, 8,
1349.
B.-Q. Xu, S-H. Chai, T. Takahashi, M. Shima, S. Sato, R. Takahashi (Nippon
Catalytic Chem. Ind.), WO 2007058221, 2007.
T. Sato, R. Takahashi (Nippon Catalytic Chem. Ind.), JP 2008088149,
2008.
H. Atia, U. Armbruster, A. Martin, J. Catal. 2008, 258, 71.
M. Okuno, E. Matsunami, T. Takahashi, H. Kasuga, M. Okada, M. Kirishik
(Nippon Catalytic Chem. Ind.), WO 2007132926, 2007.
M. Okuno, E. Matsunami, T. Takahashi, H. Kasuga (Nippon Catalytic
Chem. Ind.), JP 2007301505, 2007.
M. Okuno, E. Matsunami, T. Takahashi, H. Kasuga (Nippon Catalytic
Chem. Ind.), JP 2007301506, 2007.
X.-Z. Li (Shanghai Huayi Acrylic Acid Co), CN 101070276, 2007.
H. Jo, S.-H. Chai, T. Takahashi, M. Shima (Nippon Catalytic Chem. Ind.),
JP 2007137785, 2007.
C.-J. Zhou, C.-J. Huang, W.-G. Zhang, H.-S. Zhai, H.-L. Wu, Z. S. Chao,
Stud. Surf. Sci. Catal. 2007, 165, 527.
Q. Liu, Z. Zhang, Y. Du, J. Li, X. Yang, Catal. Lett. 2009, 127, 419.
E. Matsunami, T. Takahashi, H. Kasuga (Nippon Catalytic Chem. Ind.), JP
2007268363, 2007.
E. Matsunami, T. Takahashi, H. Kasuga (Nippon Catalytic Chem. Ind.), JP
2007268364, 2007.
H. Redlingshoefer, C. Weckbecker, K. Huthmacher, A. Doerflein (Evonik
Degussa), WO 2008092533, 2008.
H. Redlingshoefer, C. Weckbecker, K. Huthmacher, A. Doerflein (Evonik
Degussa), DE 102007004351, 2007.
H. Redlingshoefer, C. Weckbecker, K. Huthmacher, A. Doerflein (Evonik
Degussa), WO 2008092534, 2008.
H. Redlingshoefer, C. Weckbecker, K. Huthmacher, A. Doerflein (Evonik
Degussa), DE 102007004350, 2007.
L. Ning, Y. Ding, W. Chen, L. Gong, R. Lin, Y. L, Q. Xin, Chin. J. Catal.
2008, 29, 212.
W. Suprun, M. Lutecki, T. Haber, H. Papp, J. Mol. Cat. A: Chem. 2009,
309, 71.
H. Kasuga, M. Okada (Nippon Catalytic Chem. Ind.), JP 2008137950,
2008.
Y. Arita, H. Kasuga, M. Kirishiki (Nippon Catalytic Chem. Ind.), JP
2008110298, 2008.
P. OConnor, A. Corma, G. W. Huber, L. Savanaud (Bioecon. Internat.
Holding), WO 2008052993, 2008.
J.-L. Dubois (Arkema), FR 2897058, 2007.
H. Kasuga (Nippon Catalytic Chem. Ind.), JP 2008137952, 2008.
J. Kijenski, A. Migdal, O. Osawaru, E. Smigiera (Inst. Chemii Przemyslowe), EP 1860090, 2007.
J.-L. Dubois (Arkema), WO 2008129208, 2008.

Received: May 12, 2009

 2009 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

ChemSusChem 2009, 2, 719 730

S-ar putea să vă placă și