You are on page 1of 16

JID: JTICE

ARTICLE IN PRESS [m5G;August 10, 2016;13:11]

Journal of the Taiwan Institute of Chemical Engineers 0 0 0 (2016) 1–16

Contents lists available at ScienceDirect

Journal of the Taiwan Institute of Chemical Engineers


journal homepage: www.elsevier.com/locate/jtice

A review on glycerol valorization to acrolein over solid acid catalysts


Ahmad Galadima a, Oki Muraza a,b,∗
a
Center of Research Excellence in Nanotechnology, King Fahd University of Petroleum & Minerals, Dhahran 31261, Saudi Arabia
b
Chemical Engineering Department, King Fahd University of Petroleum & Minerals, Dhahran 31261, Saudi Arabia

a r t i c l e i n f o a b s t r a c t

Article history: The projected rise in global biodiesel production and demands inevitably implied a linear rise in glyc-
Received 27 March 2016 erol production. The paper therefore critically tailored recent literature on the progress made regarding
Revised 10 July 2016
the glycerol to acrolein process. Emphasis was given to the role of catalysts such as oxides, heteropoly
Accepted 20 July 2016
acids, zeolites and silicoalumino phosphates (SAPOs) and their associated activity and stability proper-
Available online xxx
ties. Concise details on the reaction mechanisms and the strategies being employed to ensure optimal
Keywords: catalytic performance, hinder deactivation and improve product yields were also appropriately captured
Glycerol and analyzed. An outlook was simultaneously tailored to identify new paths for future research.
Upgrading © 2016 Taiwan Institute of Chemical Engineers. Published by Elsevier B.V. All rights reserved.
Acrolein
Solid acids
Clean production
Catalysis

1. Introduction esters (i.e. biodiesel) and glycerol [14]. While the methyl esters
are employed as fuels, the glycerol is employed as raw-material
The unsustainability and environmental challenges associated for the petrochemical industries. The projected rise in biodiesel
with fossil fuels production and utilization have in the recent production inevitably sends a signal on the simultaneous rise in
years triggered interest for a shift to renewable feedstocks, as associated glycerol production. According to statistics, in 2008,
alternative sources of energy and petrochemicals. Burning of fossil the world production of biodiesel reached 11.1 million metric
fuels emits large quantities of CO2 and other greenhouse gases tons with 32.6 million metric tons capacity. There are projections
with strong potentials to pollute the global environment [1-5]. that by 2016, biodiesel market will hit 37 billion gallons, given
Biofuels are therefore currently considered as key prospective rise to an average growth rate of 42% per annum. According to
alternatives that could help in addressing these difficulties. Among these statistics glycerol production can hit 6 billion gallons by that
the biofuels, biodiesel is widely accepted as an important fuel period [15]. Among the key players, countries in the Asia, USA
that can be used directly or as a blend without compromising and the Western Europe stand for 75% of the net global glycerol
engine specifications [6-11]. Its production had been on the rise consumption in 2011 [16]. The Asian region consumed 36% of the
in many areas around the world. For example, according to the whole world consumption in that year. China alone accounted for
recent statistics, biodiesel production in the United States was 20 18%, with predictions that the consumption will further grow due
million gallons in 2003 but increased to 700 million gallons in to advances in the epichlorohydrin manufacture from glycerol. In
2008. In the year 2011, the production hits 1.1 billion gallons and 2016, the persistent rise in glycerol-to-pharmaceuticals in China
later 1.8 billion gallons in 2013 [12]. By the year 20 0 0, the total will shift the consumption to 27% with overall growth of 45%
global biodiesel production was 0.8 billion gallons but increased focused for the whole Asia [16]. Countries in the Western Europe
to 4 and 16 billion gallons in 2005 and 2010, respectively, with consumed about 26% in 2011 with projections of 8% growth
continuous increase predicted for the future up to 2050 [13]. annually. In other parts of Europe and Africa, the consumption had
Biodiesel is mainly produced from the process of transesterifica- been projected to rise by 7% and 9%, respectively [16]. Therefore,
tion via which fats and/or oils of plants and animals react with devising more industrial options for its valorization would be very
short chain alcohol like methanol under either homogeneous or beneficial for the industries [17-19]. Glycerol had been successfully
heterogeneous conditions. The reaction products are the methyl converted into a range of fuels and petrochemicals including diesel,
light naphtha, succinic acid, synthesis gas, olefins and polyesters

Corresponding author at: Center of Research Excellence in Nanotechnology, [14,20–28]. There are recent reports that diesel and gasoline range
King Fahd University of Petroleum & Minerals, Dhahran 31261, Saudi Arabia. hydrocarbons could be generated from glycerol [29]. Figure 1
E-mail address: omuraza@kfupm.edu.sa, oki.muraza@gmail.com (O. Muraza). illustrates some of these species that can be obtained from glyc-

http://dx.doi.org/10.1016/j.jtice.2016.07.019
1876-1070/© 2016 Taiwan Institute of Chemical Engineers. Published by Elsevier B.V. All rights reserved.

Please cite this article as: A. Galadima, O. Muraza, A review on glycerol valorization to acrolein over solid acid catalysts, Journal of the
Taiwan Institute of Chemical Engineers (2016), http://dx.doi.org/10.1016/j.jtice.2016.07.019
JID: JTICE
ARTICLE IN PRESS [m5G;August 10, 2016;13:11]

2 A. Galadima, O. Muraza / Journal of the Taiwan Institute of Chemical Engineers 000 (2016) 1–16

Fig. 1. A range of some compounds that could be derived from glycerol.

erol [30]. An important derivable industrial raw-material that is


gaining attention in the recent years is the acrolein. It is commonly
employed as good candidate for the synthesis of compounds like
amino acids, polymers, herbicides and many other classified prod-
ucts. The vast applications have thus triggered research on sus-
tainable production routes, especially in relation to catalysts and
reaction conditions [31-35]. This paper therefore critically reviewed
the recent progress on the conversion of glycerol to acrolein. Issues
related to catalysts explored and their associated activity/stability
properties were carefully and appropriately captured and
examined. Fig. 2. Chemical equation for glycerol/biodiesel production from oils/fats. Adapted
from Ref. [14].

2. Green glycerol production

Since from the 1940’s, various options have been investi- essary. These escalate the process cost and render the overall pro-
gated for commercial glycerol production. Production from propene cess less-attractive. Glycerol production via the transesterification
based intermediates such as allyl alcohol, epichlohydrin and route is also not without challenges, but is the commercial op-
propene oxide was among these early options [36,37]. With al- tion gaining much interests in the recent times, particularly due
lyl chloride, oxidation into dichlorohydrin is first achieved using to the rapid growth in the production and demands of biodiesel
hypochlorite. Hydroxide of calcium or sodium is then employed as as fuel. The transesterification process (Fig. 2) proceeds by the re-
catalyst to convert the dichlorohydrin into epichlohydrin. Glycerol action of triglyceride esters in fats/oils with short chain alcohol
is subsequently synthesized by the process of hydrolysis. During (usually methanol) to produce mono-alkyl esters (i.e. biodiesel) and
the process, aqueous NaOH or Na2 CO3 is utilized at temperatures glycerol as the only reaction products [42,43]. The reaction can
closer to 200 °C and 1 atm to achieve high yield (up to 98%) of be achieved using both homogeneous and heterogeneous catalyst
glycerol. Alternatively, propene can be transformed into propene systems [44,45]. Among the homogeneous catalysts, strong bases
oxide by the process of epoxidation, followed by subsequent iso- (e.g. hydroxides of sodium and potassium), their derived methox-
merization to produce allyl alcohol, which is subsequently epoxi- ides and common acids like hydrochloric and sulphuric acids are
dised to yield glycidol using peracetic acid. Glycerol could finally very popular [46,47].
be obtained by hydrolysis of the glycidol in basic media. The ma- Homogeneous transesterification produces low purity glycerol
jor problems with propene-intermediates-glycerol routes are; (1) because the catalyst systems are themselves soluble in both the
dependence on propene, which is mainly derived from fossil fu- biodiesel and glycerol [48]. This factor implies that, high-degree
els, as primary feedstock and (2) homogeneous nature of the reac- purification is necessary to achieve good quality product and con-
tions. These make the process unsustainable and environmentally sequently escalated process cost. Another important issue to be
challenging. considered is the corrosive nature of the hydroxides and acids, as
The two key biomass-based options (green routes) to glycerol well as their environmental disposal problems. These therefore at-
production are the hydrogenation of carbohydrates and the trans- tributed for a shift to the exploitation of a range of heterogeneous
esterification for biodiesel synthesis. Green routes to fuels and catalysts. Solid basic or acidic catalysts yield slower reaction rates
petrochemicals syntheses always target processes that ensure envi- compared to the homogeneous catalysts but they can produce
ronmental sustainability without comprising quality and efficiency equivalent yields of both biodiesel and glycerol. They can easily be
[38,39]. In the former case (i.e. carbohydrate hydrogenation), tran- separated by simple methods and are subject to recycle and reuse
sition metals based oxides usually of Ni, Cu and W, are employed given optimal performance. Catalysts in this category include metal
as catalysts to hydrogenate biomass-based polyalcohols like starch oxides and carbonates, zeolites, heteropoly acids, et cetera [49–
and sugars [40,41]. The reaction yields a mixture of glycols that 52]. Transesterification is normally conducted employing reaction
could be separated via distillation process. A considerable difficulty temperatures closer to the boiling point of the monohydric-alcohol
here is that, glycerol derived through this method is of very low (e.g. 50–65 °C for methanol), using stoichiometric oil/alcohol ra-
quality and therefore special expensive refining operations are nec- tio (i.e. 1:3). However, the process being equilibrium dependent

Please cite this article as: A. Galadima, O. Muraza, A review on glycerol valorization to acrolein over solid acid catalysts, Journal of the
Taiwan Institute of Chemical Engineers (2016), http://dx.doi.org/10.1016/j.jtice.2016.07.019
JID: JTICE
ARTICLE IN PRESS [m5G;August 10, 2016;13:11]

A. Galadima, O. Muraza / Journal of the Taiwan Institute of Chemical Engineers 000 (2016) 1–16 3

may require excess alcohol concentrations to significantly shift the fore with strong potentials to revolutionize the industries. During
process to completion [14]. In this case, effort to remove the ex- the reaction, the glycerol interacts with the acid sites (normally
cess methanol is necessary due to its associated solubility in both Brønsted) of the catalyst for subsequent dehydration to acrolein.
glycerol and biodiesel. Although the actual mechanism over solid acids is still debatable
According to a recent review by Galadima and Muraza [14], in the literature, a study by Kongpatpanich et al. [74] is in agree-
mixed or composite oxides of Mg, Ca and transition metals can ment with most findings [63,75,76]. According to this mechanism,
produce excellent yields of biodiesel (up to 98%) with associated interaction of glycerol with acid sites can proceeds through either
high purity glycerol using methanol as the monohydric alcohol at the side (i.e. primary) or the central (i.e. secondary) –OH group and
60 °C. However, sufficient reaction time like 10 h and a reactant the overall process involved three summarized steps. Initially, the
ratio (i.e. methanol to oil) of 6:1 are normally required. These secondary –OH group is removed via dehydration to yield an in-
materials could be supported over alumina or silica support to termediate specie (i.e. enol intermediate). In the subsequent step,
enhance their activities. Basic zeolites based on faujasites could a tautomerization through which the generated enol is converted
also produce comparable activity but modification with oxides into a keto specie takes place (i.e. via keto-enol tautomerism). Fi-
is usually required. Catalysts based on zirconia and organosul- nally, an –OH group from this keto intermediate gets removed to
phonic acids are challenging due to the loss of sulphate species yield acrolein. According to Yadav et al. [63], under certain reaction
and consequent acidity decay with time. However, the choice of conditions, the enol (i.e. 3-hydroxypropanal) produced can under-
appropriate dopant and zirconia phase can limit the deactivation. went cleavage to produce acetaldehyde and formaldehyde. Hydrox-
Tetragonal zirconia doped with ceria had been identified as the yacetone could also be generated when the reaction is conducted
most viable phase compared to other phases like monoclinic and at high temperature due to the dehydration of the primary –OH
cubic. Catalysts based on heteropoly acids and acidic zeolites pos- group. Oxidation of acetaldehyde to acetic acid can also be possible
sessed sufficient Brønsted acidity desired for the reaction [53-57]. (Fig. 4) [77]. Under this condition, the chance of acrolein decompo-
They are characterized by good water and free fatty acids tolerance sition into acetaldehyde and formaldehyde or CO/CO2 is also high.
and therefore resistance to deactivation. Another important issue On the basis of these observations, the catalyst acidity and reaction
is the ability to modify the structural-acidity properties of zeolites temperature are very critical factors to be considered for achieving
like H-ZSM-5, H-MOR, H-BEA etc. [58-61]. Key issues that need to desired acrolein selectivity/yield.
be further addressed are the main reaction mechanisms associated
with each category of solid acid/basic catalysts, developing simpler 4. Catalysts for glycerol to acrolein
and economical crude glycerol purification route and lowering
the time required to bring the reaction to completion with the Acrolein can successfully be synthesized from glycerol using
heterogeneous catalysts. many homogeneous and heterogeneous catalyst systems in ei-
ther the liquid of the vapor phase [78]. The dehydration can be
3. Glycerol to acrolein achieved in supercritical water to yield acrolein to good selectiv-
ity. An important benefit identified with this option is the chance
Acrolein otherwise called propenal or 2-propenal is a colorless of physical properties modification by changing pressure and tem-
to yellow liquid and the commonest unsaturated aldehyde charac- perature [74]. Watanabe et al. [79] performed glycerol conversion
terized by piercing smell, with boiling and melting points of 53 under supercritical water conditions (i.e. 400 °C and 34.5 MPa) in
and −88 °C, respectively. Acrolein can be found in cooked food the presence of H2 SO4 . They achieved 80% and 90% acrolein selec-
or in the environment [62]. Commercially, acrolein can be pro- tivity and glycerol conversion, respectively. Ott et al. [80] have re-
duced by a number of routes. Among which catalytic oxidation of ported 50% glycerol conversion with 75% acrolein selectivity at the
propene/propylene in the gaseous phase over oxides or mixed ox- temperature and pressure of 360 °C and 25 MPa, respectively, us-
ides catalysts is very popular [63]. Another important route is the ing ZnSO4 as catalyst. Other authors employed catalysts based on
propane oxidation. This route can yield both acrolein and acrylic TiO2 and WO3 , and reported complete glycerol conversion and up
acid to significant concentrations [64-66]. Catalysts based on tran- to 55% yield of acrolein [81]. However, a major drawback is poor
sition metal oxides (e.g. of vanadium and molybdenum) had been possibility of catalyst separation and the high costs required to de-
identified as good candidates for the synthesis via this route. Re- velop special quality equipment that can work under supercritical
cently, Liu and co-workers [67] reported that, propylene can se- conditions [74]. The application of appropriate catalyst can lower
lectively be oxidized into acrolein and other oxygenates like ace- the reaction temperature and possibly the process costs [30]. Het-
tone and acrylic acid using vanadium oxide catalysts supported erogeneous catalyst systems are therefore preferred. Three classes
over silica. The mechanism of the reaction was believed to involve of heterogeneous solid acid catalysts had been identified as the po-
the formation of propene oxide as a key intermediate, with subse- tential candidates for this application. They include mixed oxides,
quent formation of acrolein to high selectivity. Further oxidation of protonic zeolites and other inorganic solid acids like heteropoly
acrolein yields acrylic acid, whereas the C–C bond in acetone un- acids. The paper will therefore give emphasis to the recent stud-
derwent breakages yielding formaldehyde with possibility of fur- ies involving the solid acids.
ther oxidation into CO and CO2 species. This mechanism had been
well agreed in the recent literature [68-72]. However, the possibil- 4.1. Oxide catalysts
ity of side reactions to outweigh acrolein formation have in some
instances lower the acrolein yields. There are some other produc- Historically, the glycerol-to-acrolein production could be dated
tion routes but the former routes are the main industrial options back to the work of Schwenk et al. [82] during the 1930’s. The
largely explored [62]. Fig. 3 presents a range of possible routes authors who employed phosphates compounds of Cu and Li as
for the synthesis of acrolein, involving both the petroleum and catalysts, demonstrated that up to 80% yield of acrolein could be
biomass-based options [73]. Acrolein can be produced both via the obtained in the reaction temperature range of 30 0–60 0 °C. During
liquid and gaseous phases. that period biodiesel production was not prominent and therefore
Production of acrolein from ethane, propane or propylene relied availability of glycerol feedstock was significantly limited. For these
mainly on fossil fuels for the source of feedstock, therefore very economic reasons, glycerol dehydration to acrolein suffered a set-
challenging in terms of economic sustainability and environmen- back during the period. After many decades, a re-ignited trend of
tal considerations. Synthesis of acrolein from the glycerol is there- research continued on the subject, especially during the 1980’s and

Please cite this article as: A. Galadima, O. Muraza, A review on glycerol valorization to acrolein over solid acid catalysts, Journal of the
Taiwan Institute of Chemical Engineers (2016), http://dx.doi.org/10.1016/j.jtice.2016.07.019
JID: JTICE
ARTICLE IN PRESS [m5G;August 10, 2016;13:11]

4 A. Galadima, O. Muraza / Journal of the Taiwan Institute of Chemical Engineers 000 (2016) 1–16

Fig. 3. Routes for the production of acrolein from different feedstocks.

Fig. 4. Mechanistic approach to acrolein and side reaction products formation from glycerol using solid acid catalysts. Adapted from Ref [77].

1990’s. In the 1990’s, Neher and co-workers [83-85] demonstrated paths. However, enhanced activities were obtained when the cat-
that complete glycerol conversion with up to 75% acrolein acrolein alysts were pre-activated in n-butane/nitrogen/oxygen mixture at
selectivity could be achieved at 300°C, using acidic catalyst like 400 °C and total flowrate of 20 ml/min prior to glycerol reaction.
phosphoric acid supported over alumina. In the recent times, cat- Complete conversion (i.e. 100%) and up to 70% yield of acrolein
alysts based on oxide and heteropoly acids are receiving good at- were achieved. Similarly, acrylic acid was not produced and the
tention by researchers as glycerol dehydration to acrolein catalysts, yield of other side reaction products declined to 28.6%. The pre-
particularly due to their historical records in other related reac- treatment/pre-activation favored the creation of stronger Brønsted
tions. For instances, the heteropoly acid compounds are good can- acid sites with suitability for glycerol dehydration to acrolein. This
didates for acidic and oxidation catalysis, irrespective of whether factor prevents the reaction from progressing through the side re-
the reaction is homogeneous or heterogeneous [35]. They are char- action paths [88,89]. Catalysts based on metal pyrophosphate ox-
acterized by sufficient acidity properties with possibilities of mod- ides can form good systems for acrolein production from glycerol
ification under controlled conditions. In terms of oxides, the pure under careful activation, catalyst preparation and reaction condi-
or mixed oxides of transition metals are given preference for this tions [90-94]. The major problems with these catalysts are the de-
reaction. These materials can also possess the desired acidity prop- struction to acidity with time leading to low acrolein yields, sensi-
erties to activate glycerol and are subject to modification [78,86]. tivity to catalyst poisons and susceptibility to coke formation [95].
Recently, Feng and co-workers [87] employed vanadium phos- Chai et al. [96] conducted the reaction with catalysts based on
phate oxide catalysts synthesized with or without using polyethy- WO3 supported over a range of oxide supports (i.e. ZrO2 , Al2 O3
lene glycol during the synthesis. Their characterization data indi- and SiO2 ). Catalytic evaluations were performed at 315 °C, 400 h−1
cated the presence of both weak and medium to strong acid sites and a glycerol/water ratio of 1:9. Catalysts in which WO3 was sup-
with all the catalysts. However, the catalysts prepared by employ- ported over alumina and zirconia produced the best activity char-
ing polyethylene glycol were richer in the total acid sites. The cat- acteristics. These systems produced 70% selectivity to acrolein and
alysts were evaluated at 300 °C by employing a space velocity of yields in the range of 49–58% compared to the optimum values
4 h−1 and 36.5 wt% of glycerol in water as the feed. Catalyst syn- of 60% and 11–32% obtained with the silica supported catalysts,
thesized without polyethylene glycol produced 72% conversion of respectively. Further studies conducted with the zirconia and alu-
glycerol and 23.3% yield of acrolein, with 9.9% yield of acrylic acid. mina supported catalysts by varying the WO3 loadings from 5 to
Other side reaction products including CO and CO2 were also pro- 40 wt% and calcination temperatures from 550 to 900 °C, indicated
duced to a yield of 32.4%. When polyethylene glycol was employed 800 °C and 30 wt% as the best calcination temperature and WO3
for the synthesis, the resulting catalysts raised the conversion to loading over alumina, respectively. Catalyst designed under these
88.7% to 93.3% with 18.8–23.3% yield of acrolein and 16.1–16.7% conditions produced 70% and 61% acrolein selectivity and yield, re-
yield of acrylic acid. These catalysts also produced 46.5–54.1% of spectively. It shows the highest stability for 10 h period, produc-
other side reaction products. Therefore, the results indicated that ing a low amount of carbonaceous deposits. The overall results of
although these catalysts are richer in acid sites their strength may the study indicated the density of W species at the catalyst sur-
not necessarily be adequate for enhancing the glycerol produc- face as very critical for achieving optimal catalytic performance.
tion and consequently shifted the reaction towards the unwanted Irrespective of whether zirconia or alumina was employed as the

Please cite this article as: A. Galadima, O. Muraza, A review on glycerol valorization to acrolein over solid acid catalysts, Journal of the
Taiwan Institute of Chemical Engineers (2016), http://dx.doi.org/10.1016/j.jtice.2016.07.019
JID: JTICE
ARTICLE IN PRESS [m5G;August 10, 2016;13:11]

A. Galadima, O. Muraza / Journal of the Taiwan Institute of Chemical Engineers 000 (2016) 1–16 5

support, catalyst in which the density of W species corresponds to selectivity to acrolein. They were characterized to possess high
0.5 to 1.0 monolayer coverage yielded the best acrolein selectivity densities of acid sites (i.e. 30.0–48 mmol/g). Similarly, they were
(i.e. 69% to 72%). This parameter (i.e. surface density of W) is a fac- very stable and exhibited up to 82% glycerol conversion with con-
tor identified as very influential for reactions involving supported stant acrolein selectivity after 177 h of reaction. This stability be-
WO3 catalysts [97–100]. Under normal conditions, acrolein selec- havior was attributed to the absence of free Lewis acid sites of zir-
tivity/yield was favored by increasing the W species density up to conia that are very susceptible to coke formation. These sites have
the monolayer range (i.e. 3.5–7.6 W/nm2 ). The authors [96] further been successfully neutralized by the incorporation of niobium ox-
demonstrated that, beyond the temperature of 315 °C, the catalysts ide to form the mixed oxides. Another catalyst deactivation pos-
quickly deactivated due to production of carbonaceous species and sibility retarded was the coke depositions associated with polyg-
thus defined 315°C as the optimal temperature for these catalysts lycerols or other reaction products from acetalization of glycerol.
under the reaction parameters studied. The work of Massa and co- These types of species are generated by the catalyst basic sites and
workers [101], also employed oxides of Nb and W supported over hinder the accessibility by reacting species due to surface coverage
a range of oxide supports (i.e. TiO2 , Al2 O3 and SiO2 ) as catalysts. [107,108]. Further studies conducted by changing the temperature
The active metal oxide components were doped over the supports from 280 to 300 °C showed effect on glycerol conversion. At 280 °C,
using the incipient impregnation method. The Pyridine FT-IR re- the glycerol conversion was 64% but increased to 100% when the
sults revealed highest densities of Brønsted acid sites of 0.58 and temperature was raised to 300 °C. On other hand, the acrolein se-
0.60 μmol/m2 for the WO3 supported over alumina and titania, re- lectivity only slightly changes from 71% to 70%. Among other side
spectively, than for all other catalysts. Glycerol activations were reaction products, hydroxyacetone dominated with 16.6% and 11.8%
conducted employing 0.8 g of catalyst, 20 wt% of glycerol in water selectivities at 280 and 300 °C, respectively. CO2 was not detected
and 0.94 h−1 space velocity at 305 °C. The activity was mainly de- at all the temperatures but CO was produced only to a maximum
pendent on the catalyst/support composition. Reactions conducted selectivity of 2.2%.
with 1 wt% each of NbOx and WO3 supported over alumina pro- Recently, Ginjupalli and co-workers [109] investigated the role
duced 97.1% and 98.6% glycerol conversion, respectively. The re- of zirconia structure for 10 wt% WO3 supported catalysts. Accord-
spective acrolein selectivities were 59.9% and 67.0%. When the re- ing to their characterization data, the pure forms of the tetragonal
actions were repeated with silica as the support the respective and monoclinic zirconia supports possessed 1.20 and 1.82 μmol/m2
conversions reduced to 82% and 28.2%, whereas the acrolein se- acid sites densities with associated Brønsted/Lewis (B/L) acid sites
lectivity to 57.2% and 49.3%. Incorporation of titania as support ratios of 0.307 and 0.975, respectively. When loaded with 10 wt%
yielded 98.9% glycerol conversion and 55.4% and 66.2% acrolein se- WO3 , the respective acid site densities increased to 4.39 and
lectivity for the 1 wt% NbOx and 1 wt% WO3 supported catalysts, 6.38 μmol/m2 . Similarly, the respective B/L ratios were 1.428 and
respectively. According to these results, titania formed the best 2.250. The results thus indicated the monoclinic phase as more
support followed by alumina, with the least performance observed susceptible for the Brønsted acid sites generation. Glycerol conver-
with the catalysts supported over silica. Further studies conducted sion and acrolein selectivity correlated well with this acidity trend.
with supported NbOx -WO3 composite catalysts, indicated 0.25 wt% For the reactions conducted at 300 °C using 0.2 g of catalyst and
NbOx -0.75 wt%WO3 supported over titania as the best catalyst, pro- 20 wt% glycerol in water, the pure forms of tetragonal and mono-
ducing 98.8% and 65.2% glycerol conversion and acrolein selec- clinic zirconia produced respective glycerol conversions of 65% and
tivity, respectively. The authors [101] attribute the highest activi- 70% and acrolein selectivities of 15% and 22%. When the reactions
ties achieved with the TiO2 and Al2 O3 supported catalysts to the were repeated with the 10 wt% WO3 loaded catalysts, the conver-
highest densities of Brønsted acid sites compared to the SiO2 sup- sion increased to 100%, whereas the acrolein selectivities to 70%
ported catalysts, as also demonstrated elsewhere that [102–104], and 75%, respectively. The nature of the results provides an impor-
these acid sites correlated well with acrolein selectivity. In a re- tant insight on the mechanism of the reaction involving the solid
lated study, Massa et al [105] employed ZrO2 as the support. Reac- acid catalysts. Two vital routes were considered; (1) reaction in-
tions were similarly conducted using 0.8 g of catalyst and 20 wt% volving the Lewis acid sites to yield mainly hydroxyacetone and
of glycerol in water at 305 °C. Without any modification, the pure (2) reaction involving the Brønsted acid sites to produce acrolein
zirconia support produced 85.7% conversion with only 7.1% selec- as the dominant product to highest selectivity. These observations
tivity to acrolein. Catalyst containing 1 wt% of Nb oxide supported correlated very well with some earlier proposals [76,110-112]. It
over zirconia produced 99.6% glycerol conversion and 75.1% selec- could be established that, poor selectivity to acrolein of the pure
tivity to acrolein. When the experiment was repeated with 1 wt% zirconia supports was due to low B/L ratio. On the contrary, the
W oxide supported over zirconia, the conversion slightly reduced high B/L ratios associated with the 10 wt% WO3 doped catalysts
to 98.3% but the selectivity increased 76.5%. According to these re- shifted the reaction pathway towards enhanced acrolein selectiv-
sults, zirconia formed the best support compared to silica, alumina ity consequent to the ability of Brønsted acid sites to actively and
or titania employed by the authors in the previous study [101]. sufficiently dehydrate glycerol.
In a similar trend, the overall results indicated catalyst based on Guan and co-workers [113], explored the role of catalyst syn-
tungsten oxide as the best catalyst, producing 78% yield of acrolein thesis procedure on the activity of zirconium phosphates (i.e.
compared to 75% obtained with the niobium oxide based catalyst. P/ZrO2 ) catalysts. Three synthesis methods; hydrothermal, precip-
Correlation with characterization results indicated Brønsted acid itation and impregnation were adopted. The catalyst acidity was
sites as the necessary active sites desired for the reaction to pro- dependent on the preparation option employed. Catalysts prepared
ceeds through the glycerol formation pathway and that, catalysts by hydrothermal and impregnation methods possessed the acid
richer in their densities are more favorable. Thus, incorporation of sites densities of 46 and 47 μmol/m2 , whereas the catalyst pre-
zirconia as support under constant conditions enhances the gener- pared by the precipitation method (calcined at 100o ) possessed
ation of the active Brønsted acid sites than when other supports 13 μmol/m2 . When reactions were conducted using 0.5 g of cata-
like silica, alumina or titania were incorporated. lyst at 315 °C employing 10 wt% of glycerol in water, the respec-
The effect of catalyst stability was investigated by Lauriol- tive conversions were 52%, 63.3% and 100% for the catalysts devel-
Garbay and co-workers [106]. They employed mixed niobium and oped via impregnation, hydrothermal and precipitation procedures.
zirconium oxide catalysts (i.e ZrNbO) under the reaction conditions On the other hand, the respective acrolein selectivities were 18.1%,
of 300 °C and 20 wt% glycerol in water with a flowrate of 3.8 g h−1 . 33.9% and 70.9%. The lower activities observed with the catalysts
The catalysts produced 100% glycerol conversion with up to 72% developed through the first two methods have been attributed to

Please cite this article as: A. Galadima, O. Muraza, A review on glycerol valorization to acrolein over solid acid catalysts, Journal of the
Taiwan Institute of Chemical Engineers (2016), http://dx.doi.org/10.1016/j.jtice.2016.07.019
JID: JTICE
ARTICLE IN PRESS [m5G;August 10, 2016;13:11]

6 A. Galadima, O. Muraza / Journal of the Taiwan Institute of Chemical Engineers 000 (2016) 1–16

their high concentrations of weak acid sites. Therefore, both the


number and the strength of acid sites play a vital role during glyc-

[115]
[115]
[115]
[114]
[114]
[114]

[116]
[116]

[118]
[118]
[118]
[117]
[117]
Ref.
erol to acrolein conversion. Further studies performed to establish
the effect of catalyst calcination temperature with the catalysts de-
veloped via the precipitation method showed the acrolein selec-

Catalyst Acidity, mmol/g


tivity to increase from 70.9% to 81.5% at a constant conversion of
100% for the calcination temperatures of 100 and 400 °C, respec-
tively. However, beyond 400 °C the conversion was stable, but the

Not reported
Not reported
Not reported
Not reported
Not reported
selectivity declined. For instance, at 500 °C the selectivity reduced
to 72.8% and when the calcination was conducted at 700 °C the
selectivity declined to 63.4%. At these higher calcination conditions

0.66
0.66
3.84
3.72

4.91
1.87
1.54
1.10
the strong acid sites of the zirconium phosphate catalysts were sig-
nificantly altered.
In Table 1, the roles of oxide catalysts modifications on catalytic

Acrolein selectivity, %

40, increased to 45
activity were tailored. According to Tao et al. [114], the nature of

35, reduced to 15
a mixed oxide catalyst can affect its acidity and catalytic prop-
erties. Mixed oxide of 50 wt% TiO2 –SiO2 possessed a total acid-

40, stable
ity of 1.10 mmol/g compared to 1.54 mmol/g obtained when SiO2
was replaced with ZrO2 to form 50 wt% TiO2 –ZrO2 catalyst sys-

62.6
62.1
62
82

65
70
57

10
10
10
tem. However, a mixed oxide catalyst containing 10 wt% ZrO2 –SiO2
showed the highest acidity of 1.87 mmol/g compared to these cat-

Glycerol conversion, %
alysts. When glycerol conversions were conducted at 315 °C using

100, reduced to 75
100, reduced to 95
10 mol% of glycerol in water for 10 h, glycerol conversion, acrolein

78, reduced to 60
75, reduced to 35
30, reduced to 12
selectivity and catalyst lifetime were all dependent on the mixed
oxide catalyst nature. The glycerol conversion for 50 wt% TiO2 –
SiO2, 10 wt% ZrO2 –SiO2 and 50 wt% TiO2 –ZrO2 were 30%, 75% and
78%, respectively. During the 10 h period, the values dropped to

89.2
65.5
100
100

95

58
90
97
12%, 35% and 60%, respectively. On the other hand, the respec-
tive acrolein selectivities were 40%, 35% and 40%. For the 50 wt%

Glycerol/water ratio of 2.3:46.3 in nitrogen ratio of 51.4, 300 °C, 2900 h−1 , 80 h.
Glycerol/water ratio of 2.3:46.3 in nitrogen ratio of 51.4, 300 °C, 2900 h−1 , 80 h.
TiO2 –SiO2 catalyst, the selectivity was completely stable for 10 h

0.1 g of catalyst, 360 °C, 18, 0 0 0 h−1 , glycerol/oxygen/water/helium = 2:4:40:54.


0.1 g of catalyst, 360 °C, 18, 0 0 0 h−1 , glycerol/oxygen/water/helium = 2:4:40:54.
0.1 g of catalyst, 360 °C, 18, 0 0 0 h−1 , glycerol/oxygen/water/helium = 2:4:40:54.
whereas the selectivity dropped from 35% to only 15% for the

10 mol% glycerol in water, 400 h−1 , 315 o C, 10 h. Catalyst calcined at 500 o C


10 mol% glycerol in water, 400 h , 315 °C, 10 h. Catalyst calcined at 700 C

10 mol% glycerol in water, 400 h−1 , 315 °C, 10 h. Catalyst calcined at 550 o C
o
10 wt% ZrO2 –SiO2 catalyst. On the contrary, the acrolein selectivity
of 50 wt% TiO2 –ZrO2 catalyst increased from 40% to 45%. According
to the overall results, moderate concentration of acid sites yielded

20 wt% of glycerol, 340 °C, 873 h−1 , 3 h, catalyst calcined 550 °C.
20 wt% of glycerol, 340 °C, 873 h−1 , 3 h, catalyst calcined 550 °C.
0.2 g of catalyst, 20 wt% glycerol, 300 °C and nitrogen/air of 7:3.
0.2 g of catalyst, 20 wt% glycerol, 300 °C and nitrogen/air of 7:3.
the best activity with these catalysts. At high concentration, the 0.2 g of catalyst, 20 wt% glycerol, 300o C and nitrogen/air of 7:3.
acid sites are prone to cause the coke formation and consequently
lower the catalyst lifetime. Srinivasa and co-workers [115] studied
the effect of WOx loading onto phosphate modified ZrO2 (i.e. P–
ZrO2 ). Increasing WOx loadings from 5 to 30 wt% increases cata-
lyst acidity from 3.72 to 4.91 mmol/g. When glycerol conversions
were performed at 300 °C using 20 wt% of glycerol in water and
Activities of some modified oxide catalysts during glycerol upgrading into acrolein.

air/nitrogen ratio of 3:7, both glycerol conversion and acrolein se-


−1

lectivity followed a similar trend. The glycerol conversions were


97%, 100% and 100% for the WOx loadings of 5, 10 and 30 wt%, re-
spectively. Similarly, the respective acrolein selectivity were 57%,
62% and 82% for these respective loadings. Therefore, in this case,
high acidity was more favorable for acrolein production. The role
Reaction conditions

of using air or nitrogen as carrier gas can also influence the cat-
alytic activity as demonstrated by Gu et al. [116]. Reactions car-
ried out with a 29 wt% NiSO4 /SiO2 catalyst at 340 °C and 873 h−1
for 3 h, indicated a decreased in glycerol conversion from 89.2%
when the reaction was conducted in air to 65.5% when nitrogen
was employed as the carrier gas. However, both conditions pro-
duced comparable acrolein selectivity. Therefore, incorporation of
29 wt%NiSO4 /SiO2 (Nitrogen as carrier)

air into the feed appeared more favorable with this catalyst system.
29 wt%NiSO4 /SiO2 (Air as carrier)

According to Znaiguia and co-workers [117] a slight change in cat-


10 wt% ZrO2 –SiO2 , mixed oxide

2.4 wt% WO3 /0.6 wt% SiO2 /ZrO2


2.3 wt% WO3 /0.5 wt% SiO2 /ZrO2
50 wt% SiO2 –TiO2 , mixed oxide

50 wt%TiO2 –ZrO2, mixed oxide

alyst composition can cause significant changes to the catalyst life-


time and acrolein selectivity for WO3 –SiO2 /ZrO2 catalysts. When
glycerol conversions were conducted with 2.4 wt% WO3 -0.6 wt%
30 wt% WOx /P-ZrO2
10 wt% WOx /P-ZrO2

3.8 wt%SO4 2− /ZrO2


4.4 wt%SO4 2− /ZrO2
5 wt% WOx /P-ZrO2

SiO2 /ZrO2 , complete (i.e. 100%) glycerol conversion that reduced


2 wt%SO4 2− /ZrO2

to 75% and acrolein selectivity of 65% were achieved after 80 h.


However, better results were obtained when the catalyst composi-
tion was slightly modified to 2.3 wt% WO3 -0.5 wt% SiO2 /ZrO2 under
Catalyst
Table 1

constant reaction conditions (i.e. 300 °C, 2900 h−1 and 80 h). Glyc-
erol conversion slightly reduced only from 100% to 95% whereas
the acrolein selectivity rose to 70%. According to a study by Ca-

Please cite this article as: A. Galadima, O. Muraza, A review on glycerol valorization to acrolein over solid acid catalysts, Journal of the
Taiwan Institute of Chemical Engineers (2016), http://dx.doi.org/10.1016/j.jtice.2016.07.019
JID: JTICE
ARTICLE IN PRESS [m5G;August 10, 2016;13:11]

A. Galadima, O. Muraza / Journal of the Taiwan Institute of Chemical Engineers 000 (2016) 1–16 7

vani et al. [118], incorporation of SO4 2− species can influence the ter/steam tolerance of this catalyst. On the other hand, the low
activity of ZrO2 at 360 °C, 18,0 0 0 h−1 and a glycerol/water ratio of catalytic performance of the HPW and HPMo catalysts was a con-
2:40 reaction conditions. Increasing the SO4 2− from 2 to 4.4 wt% sequence of low Brønsted acidity and poor stability and water-
reduced the glycerol conversion from 95% to 58% with very low tolerance as also observed elsewhere [129]. Regarding the acrolein
unaltered acrolein selectivity of 10%. The low yield of acrolein with yields, a similar trend was observed with glycerol conversion.
these catalysts have generally been attributed to the loss of SO4 2- HSiW produced the highest acrolein yield of 78.6% compared to
species during the reaction and associated coke formation. The the maximum yields of 68.7% and 55.9% obtained with the HPW
presence of these species is necessary to derive sufficient Brønsted and HPMo catalysts, respectively. Shen et al. [130] incorporated
acidity desired to produce high yields of acrolein. Acrolein synthe- TiO2 as a support material for the same heteropoly acids (i.e. HSiW,
sis from glycerol over oxide catalysts is therefore dependent not HPW and HPMo). Catalyst loadings over the TiO2 support were
only on the catalyst acidity but other parameters such as catalyst varied from 5 to 20 wt%. The low acidity (0.26 mmol/g) of the
textural properties, preparation conditions, influence of air or ni- TiO2 generally increased following the incorporation of heteropoly
trogen as the carrier, reaction temperature and glycerol feed con- acids but the modification was dependent on the catalyst nature.
centration as also demonstrated elsewhere [119-122]. The highest total acidity of 3.0 mmol/g was observed with the
A series of recent studies conducted on the glycerol valoriza- 20 wt% HSiW/TiO2 whereas the least total acidity of 0.68 mmol/g
tion to acrolein with oxides focused on developing new strate- was observed with the 5 wt% HPMo/TiO2 catalyst. Glycerol con-
gies for improving acrolein selectivity, reducing the production of versions were conducted in the temperature range of 220–300 °C.
by-products and hindering catalyst deactivation by carbobnaceous Although all the catalysts produced reasonably high glycerol con-
depositions [123,124]. Dalil et al. [125], demonstrated that at the versions that increased with increase in temperature, the activity
temperature of 280 °C, WO3 /TiO2 can produce 73% selectivity to was temperature and catalyst dependent. For instance, with 20 wt%
acrolein within 6 h reaction period. The production of by-products HSiW/TiO2 , 20 wt% HPW/TiO2 and 20 wt% HPMo/TiO2 catalysts, the
(i.e. propanal, acetaldehyde, formic and acetic acids) declined by conversion increased from 95.9% to 100%, 88.7% to 93.9% and 65%
40% within similar period. They attributed partial coke-oxidation to 95.3% when the temperature was raised from 220 to 300 °C
as responsible for the reduction in carbonaceous deposition from for each catalyst, respectively. Catalysts based on HSiW/TiO2 pro-
34% to only 6%. It could be established that the synergic interaction duced the highest conversions at all the temperatures and compa-
between WO3 and TiO2 support as well as the dispersion of the rable loadings whereas the least conversions were observed with
WO3 particles are very essential for improving catalytic activity- HPMo/TiO2 based catalysts. Supporting HSiW with TiO2 enhanced
stability properties. Cecilia et al. [126], showed the incorporation the creation of stronger Brønsted acid sites, water tolerance and
of mesoporous SBA-15 doped Zr as a support to be a good option stability than for the other catalysts. The trend was similar for
for improving the catalytic activity of the WO3 material. The sup- the acrolein selectivity. Catalysts based on HSiW/TiO2 produced
ported catalyst produced 97% glycerol conversion and 41% yield of the highest acrolein selectivity of 80.0% compared to the maxi-
acrolein compared to less than 30% achieved with the unsupported mum selectivities of 49.6% and 42.0% achieved with the HPW/TiO2
catalyst. The mesoporosity of Zr doped SBA-15 support enhanced and HPMo/TiO2 based catalysts, respectively. The overall catalytic
the diffusion of both glycerol and acrolein and consequently lim- results provide an insight into the reaction mechanism involving
ited the occurrence of side reactions. these systems as also proposed elsewhere [77,112,131-133]. The ini-
tial steps proceed by glycerol protonation involving the secondary
–OH group by the Brønsted acid sites of the heteropoly acid fol-
4.2. Heteropoly acid catalysts lowed by subsequent dehydration to 3-hydroxypropanal. Dehydra-
tion of the 3-hydroxypropanal yields acrolein. Therefore, the ability
Compounds of heteropoly acids containing reasonable densi- of HSiW/TiO2 catalysts to produce the best catalytic activity could
ties of strong Brønsted acid sites are also employed for glycerol be explained by its Brønsted acidity.
conversion. The heteropoly acids are usually modified by incor- Recently, Liu and co-workers [134] investigated the role of calci-
poration of metals like Cs to yield corresponding acidic salts or nation temperature on the activities of Al2 O3 supported HSiW (i.e.
supported over oxides such as ZrO2 , SiO2 , Al2 O3, TiO2 etc to en- HSiW/Al2 O3 ) synthesized via the impregnation method. The cata-
sure improved catalytic performance. Shen et al. [127] conducted lysts were calcined in the temperature range of 350 to 650 °C for
a study with a range of heteropoly acids; silicotungstinic acid a period of 4 h. Catalytic evaluations were performed using the
(H4 SiW12 O40, HSiW), phosphotungstinic acid (H3 PW12 O40, HPW) conditions of 300 °C, 20 wt% of glycerol in water (i.e. 1.5 mL/h).
and phosphomolybdic acid (H3 PMo12 O40, HPMo) catalysts. The The total acidity and B/L ratio were all dependent on the calci-
catalysts were characterized by total acidity of 5.40, 4.02 and nation temperature. Al2 O3 support possessed a low B/L ratio of
3.52 mol.NH3 /mol.cat for the HPW, HSiW and HPMo, respectively. 0.03 and total acidity of 0.014 mmol/g that increased with HSiW
When glycerol conversions were carried out at 320 °C, using 1 loading. The highest acidity of 0.87 mmol/g and B/L ratio of 0.78
mole of glycerol and variable catalyst loadings (0.1 to 2.1 mmol), were observed with the uncalcined HSiW/Al2 O3 . These values de-
the glycerol conversion reached 100%, but the time taken to reach creased linearly when the calicination temperature was raised to
this complete conversion was much dependent on the catalyst 550 °C. At 350 °C calcination temperature, the respective B/L ratio
loading and its nature. Reactions with 0.1, 0.2 and 0.3 mmol of and total acidity were 0.67 and 0.674 mmol/g but decreased to 0.33
HSiW reached completion after 40, 30 and 20 h, respectively. When and 0.362 mmol/g when the calcination temperature was raised
the reactions were repeated with 0.2, 0.3 and 0.4 mmol of HPW, to 550 °C. Further raising the calcination to 650°C however, raised
complete methanol conversion was achieved after 30, 20 and 20 h, the B/L ratio to 0.56 with a further decreased in the total acidity
respectively. For the respective HPMo catalyst loadings of 0.9, 1.5 to 0.319 mmol/g. Consistent with previous studies [102, 135-137],
and 2.1 mmol, complete conversion was achieved after 50, 50 and both Brønsted and Lewis acid sites were detected. All the cata-
30 h. The overall results indicated the catalytic activity to follow lysts showed an increased in glycerol conversion when the space
the order HSiW > HPW > HPMo. The best activities (i.e. equiva- velocity was raised from 0.6 to 1.0 h−1 . The conversion of Al2 O3,
lent conversion at lower catalyst loading) observed with the HSiW uncalcined HSiW/Al2 O3 and HSiW/Al2 O3 calcined at 350, 450, 550
catalyst could be attributed to higher catalyst stability and posses- and 650 °C increased from 44.3% to 47.4%, 85.8% to 100%, 80.9% to
sion of higher strong Brønsted acid sites desired for the reaction 100%, 78.9% to 100%, 65.7% to 100% and 75.5% to 100%, respectively.
[77,128]. Another important factor could be due the higher wa- A similar trend could be observed with the yields of glycerol. The

Please cite this article as: A. Galadima, O. Muraza, A review on glycerol valorization to acrolein over solid acid catalysts, Journal of the
Taiwan Institute of Chemical Engineers (2016), http://dx.doi.org/10.1016/j.jtice.2016.07.019
JID: JTICE
ARTICLE IN PRESS [m5G;August 10, 2016;13:11]

8 A. Galadima, O. Muraza / Journal of the Taiwan Institute of Chemical Engineers 000 (2016) 1–16

respective yields of acrolein were 15.0% to 19.2%, 63.4% to 66.0%,


64.1% to 64.5%, 55.9% to 57.1%, 53.1% to 63.5% and 55.7% to 58.7%.

[144]
[144]
[144]
[144]
[142]
[142]
[142]
[142]
[143]
[143]
[143]

[145]
[145]
[145]
[145]
[77]
[77]
[77]
Ref.
The overall catalytic results indicated the reaction to be acidity de-
pendent. The uncalcined catalyst produced the highest activity due
to associated highest acidity and B/L ratio. Similarly, low calcina-

Acrolein selectivity, %
tion temperatures that yield high acidities produced high activities
under constant conditions. However, stability studies conducted in-
dicated the HSiW/Al2 O3 calcined at 650 °C as more stable than the
most active uncalcined HSiW/Al2 O3 catalyst for a 7 h period. Al-
though acrolein yield was stable with both catalysts, the glycerol

66.7

66.2
23.3

25.7
34.3

33.4
75.1

65.1

74.0
27.3
conversion for the uncalcined HSiW/Al2 O3 catalyst reduced from

0.0
94
75

96
93
80

90
72
100% to 75% compared to 100% to 92% for the HSiW/Al2 O3 cata-
lyst calcined at 650 °C. The increase in activity with B/L ratio for

Conversion, %
these catalysts showed the Brønsted acid sites as the most impor-
tant active sites for acrolein formation as also proposed previously
[138-140]. In a related development, Balaga et al. [141] employed

98.4
22.9

26.6

92.6

99.7
72.6

76.6
13.2
15.7

16.4

100

100

100
heteropoly acids containing V and Mo supported over SBA-15 that

98
98

82
97

97
was mainly mesoporous in nature. The loading of the heteropoly
acid (H4 PMo11 VO40 ) was varied from 10 to 50 wt%. The character-

0.75 (0.05 Brønsted)

1.03 (0.44 Brønsted)


0.78 (0.16 Brønsted)

0.19 (0.10 Brønsted)


ization data showed the catalyst Brønsted and total acidity to in-
crease when the loading of the heteropoly acid was increased. The

Acidity, mmol/g

Not quantified
Not quantified
Not quantified
Not quantified
Not quantified
Not quantified
Not quantified
Not quantified
Not quantified
Not quantified
Not quantified
Not quantified
Not quantified
Not quantified
various loadings also retained the mesoporous properties of the
SBA-15 support. When the reactions were carried out using 10 wt%
of glycerol in water in the gas phase, the catalysts exhibited 100%
conversion but the selectivity to acrolein was higher for the cata-
lysts possessing higher densities of Brønsted acid sites. The activity
was also supported by high dispersion of the heteropoly acid over

Catalysts calcined at 500 °C, 0.5 g of catalyst, 0.24 h−1 , 300 °C, 20 wt% glycerol in water, 2 h.
Catalysts calcined at 500 °C, 0.5 g of catalyst, 0.24 h−1 , 300 °C, 20 wt% glycerol in water, 2 h.
Catalysts calcined at 500 °C, 0.5 g of catalyst, 0.24 h−1 , 300 °C, 20 wt% glycerol in water, 2 h.
the SBA-15 support as observed from the X-ray diffraction data.
In Table 2, the activities of some modified or supported
heteropoly acids recently reported in the literature have been
documented. According to Kang and co-workers [142], incor-
Some recent studies on the behaviors of heteropoly acids during acrolein synthesis from glycerol. AC = Activated carbon.

Glycerol to water ratio of 1:46, 0.38 to 0.48 g of catalyst, 3.8 h−1 , 330 °C, 5 h.
Glycerol to water ratio of 1:46, 0.38 to 0.48 g of catalyst, 3.8 h−1 , 330 °C, 5 h.
Glycerol to water ratio of 1:46, 0.38 to 0.48 g of catalyst, 3.8 h−1 , 330 °C, 5 h.
Glycerol to water ratio of 1:46, 0.38 to 0.48 g of catalyst, 3.8 h−1 , 330 °C, 5 h.
poration of mixed SiO2 and Al2 O3 as support can modify the
acidity and catalytic properties of H3 PW12 O40 (HPW) catalyst.
Catalyst supported over SiO2 and Al2 O3 as different supports (i.e.
HPW/SiO2 and HPW/Al2 O3 ) possessed low Brønsted acidity of 0.10
and 0.05 mmol/g, respectively. However, the Brønsted acidity was
promoted following the incorporation of SiO2 into the Al2 O3 sup-
0.1 g of catalyst, 7 mol% of glycerol in water, 275 °C, 10 h.
0.1 g of catalyst, 7 mol% of glycerol in water, 275 °C, 10 h.
0.1 g of catalyst, 7 mol% of glycerol in water, 275 °C, 10 h.
0.1 g of catalyst, 7 mol% of glycerol in water, 275 °C, 10 h.

ported catalysts. For the two catalysts HPW/15 mol% SiO2 –Al2 O3
250 °C, 24,0 0 0 h−1 , 20 wt% of glycerol in water, 3 h.
250 °C, 24,0 0 0 h−1 , 20 wt% of glycerol in water, 3 h.
250 °C, 24,0 0 0 h−1 , 20 wt% of glycerol in water, 3 h.
250 °C, 24,0 0 0 h−1 , 20 wt% of glycerol in water, 3 h.
and HPW/50 mol% SiO2 –Al2 O3 , the Brønsted acidity increased to
0.16 and 0.44 mmol/g, respectively. Glycerol conversions conducted

0.3 g of catalyst, 1.8 mol/h of glycerol, 325 o C


0.3 g of catalyst, 1.8 mol/h of glycerol, 325 o C
0.3 g of catalyst, 1.8 mol/h of glycerol, 325 o C
at 275 °C using 7 mol% of glycerol in water for a period of 10 h
showed the reaction to be dependent on the Brønsted acidity.
The separate SiO2 and Al2 O3 supported catalysts produced respec-
tive glycerol conversions of 16.4% and 13.2% and corresponding
acrolein selectivities of 25.7% and 23.3%. On the other hand,
the catalysts containing 15 and 50 mol% of SiO2 produced the
respective conversions of 15.7% and 22.9% and corresponding
acrolein selectivities of 27.3% and 34.3%. The results thus indicated
Reaction conditions

mixed oxide supported catalysts as the most active due to the


higher Brønsted acidity properties. Liu et al. [143] conducted the
reaction with variable loadings of Cs modified heteropoly acids
(i.e. Cs2.5 H0.5 PW12 O40 , Cs-HPW) supported over Nb2 O5 calcined
at 500 °C. Glycerol conversions were performed at 300 °C using
20 wt% of glycerol in water and 0.24 h−1 for a period of 2 h.
Catalysts containing the heteropoly acid loadings of 5 and 20 wt%
20 wt.% Cs2.5 H0.5 PW12 O40 / Nb2 O5
40 wt% Cs2.5 H0.5 PW12 O40 / Nb2 O5

(i.e. 5 wt% Cs-HPW/ Nb2 O5 and 20 wt% Cs-HPW/ Nb2 O5 ) produced


5 wt% Cs2.5 H0.5 PW12 O40 / Nb2 O5
H3 PW12 O40 /50 mol%SiO2 –Al2 O3
H3 PW12 O40 /15 mol%SiO2 –Al2 O3

similar glycerol conversion of 98%, but the acrolein selectivity


increased from 72% for the 5 wt% loading to 75% for the 20 wt%
30 wt% H4 PMo12 O40 /SiO2
0.05 mol%Rb/ HSiW12 O40
0.25 mol%Rb/ HSiW12 O40

30 wt.% HSiW12 O40 /SiO2


0.05 mol%Cs/ HSiW12 O40
0.25 mol%Cs/ HSiW12 O40

30 wt% H3 PW12 O40 /SiO2


10 wt % HSiW12 O40 /AC
30 wt% HSiW12 O40 /AC

loading. When the heteropoly acid loading was raised to 40 wt%


5 wt% HSiW12 O40 /AC

(i.e. to form 40 wt% Cs-HPW/ Nb2 O5 catalyst), the conversion


H3 PW12 O40 /Al2 O3

H3 PW12 O40 /SiO2

slightly reduced to 97%, whereas the selectivity increased to 80%.


Therefore, the higher loadings yield positive effect on acrolein
HSiW12 O40

yields. Similarly, these Cs containing catalysts were very stable


Catalyst
Table 2

for the 2 h reaction period. Haider et al. [144] employed 0.05 and
0.25 mol% each of Cs and Rb as modifiers for HSiW12 O40 (HSW)
catalyst. Reactions performed at 250 °C and 24,0 0 0 h−1 using

Please cite this article as: A. Galadima, O. Muraza, A review on glycerol valorization to acrolein over solid acid catalysts, Journal of the
Taiwan Institute of Chemical Engineers (2016), http://dx.doi.org/10.1016/j.jtice.2016.07.019
JID: JTICE
ARTICLE IN PRESS [m5G;August 10, 2016;13:11]

A. Galadima, O. Muraza / Journal of the Taiwan Institute of Chemical Engineers 000 (2016) 1–16 9

20 wt% of glycerol in water, showed both glycerol conversion and (i.e. Brønsted or Lewis) should be considered most appropriate for
acrolein selectivity to decrease with increased in metal mol%. The these catalysts. With catalysts based on oxides, Brønsted acid sites
glycerol conversions for 0.05 mol% Cs/HSW and 0.05 mol% Rb/HSW are considered the main active centers required for acrolein for-
were both 100% and the corresponding selectivities were 96% and mation. However, a recent study involving zeolites by Wang et al.
94%, respectively. However, these values decreased to 97% and 82% [171] suggested the participation of both the Brønsted and Lewis
for conversions and 93% and 90% for selectivities when the reac- acid sites. The authors proposed that, the secondary –OH group of
tion was repeated with 0.25 mol% Cs/HSW and 0.25 mol%Rb/HSW, glycerol is first dehydrogenated over the Brønsted acid sites fol-
respectively. Among the catalysts, 0.05 mol% Cs/HSW yields the lowed by subsequent dehydration involving the Lewis acid sites to
best catalytic activity (i.e. 100% conversion and 96% selectivity), produce acrolein. The study was conducted with a range of ZSM-
indicating low Cs concentration as a more appropriate modifica- 5 zeolite catalysts of varying acidity properties. Al/H-ZSM-5 pos-
tion. Ning and co-workers [145], employed various loadings of the sessed the highest acidity of 0.956 mmol/g, whereas Al/Na-ZSM-5
heteropoly acid HSW supported over activated carbon (AC) for have the least acidity of 0.434 mmol/g. When glycerol conversion
acrolein production at 330°C, using glycerol to water ratio of 1:46 was conducted at 315 °C using 50 mg of catalyst and 36.2 wt% of
for 5 h and 3.8 h−1 space velocity. Increasing the HSW loading glycerol in water, the glycerol conversion increased with increasing
from 5 to 10 wt% increases the glycerol conversion from 72.6% acidity. For the least acidic Al/Na-ZSM-5 catalyst, the glycerol con-
to 92.6% and acrolein selectivity from 66.7% to 75.1%. However, version was 36% but increased to 46% and 85% when the reaction
a further increased in the loading to 30 wt% yields a negative was repeated with H-ZSM-5 and Al/H-ZSM-5, respectively. A simi-
effect. The glycerol conversion and acrolein selectivity reduced to lar trend was observed with the acrolein yield. With Al/Na-ZSM-5,
76.6% and 66.2%, respectively. According to these results, 10 wt% the yield was 22.7% but increased to 35.9% and 54.4% for the H-
HSW/AC formed the optimal catalyst. Therefore, in addition to ZSM-5 and Al/H-ZSM-5, respectively. In correlation with previous
catalyst acidity, careful selection of active component loading and literature, incorporation of Al species into the zeolites increases
a support combination are necessary to derive the reaction to- the density of Lewis acid sites at the surface and promotes the
wards optimal acrolein yields. Tsukuda et al. [77] investigated the strength of the active Brønsted acid sites [172-176]. This factor fa-
role of heteropoly acid nature on the activities of SiO2 supported vorably enhanced glycerol activation to acrolein.
catalysts. 30 wt% each of H3 PW12 O40 (HPW), H4 SiWO40 (HSW) Dalla and co-workers [177] evaluated the role of La modification
and H4 PMo12 O40 (HPMo) was loaded a over commercial grade on the activity of H-BEA (i.e. Beta zeolite) catalyst. The total acidity
SiO2 . Glycerol activation was conducted at 325 °C, using 0.3 g of and B/L ratio were dependent on this modification. The respective
catalyst and 1.8 mol/h feed flow of glycerol. The 30 wt% HSW/SiO2 acidities of the parent H-BEA and the two La modified catalysts
produced the highest glycerol conversion of 100% but the least (i.e La/H-BEA-1 and La/H-BEA-2) were 1.08, 1.00 and 0.66 mmol/g.
acrolein selectivity of 33.4%. With 30 wt% HPW/SiO2 and 30 wt% The corresponding B/L acidity ratios were 1.86, 2.09 and 4.05, re-
HPMo/SiO2, the conversion slightly reduced to 99.7% and 98.5% spectively. The results indicated a decreasing acidity and increas-
whereas the selectivity increased to 65.1% and 74.0%, respectively. ing the B/L ratio with the incorporation of La species. Activation
These catalysts exhibited stronger Brønsted acidity suitable for of glycerol performed at 275 °C and 0.75 h−1 for 7 h showed the
acrolein formation than the 30 wt% HSW/SiO2 catalyst. Glycerol catalysts to produce high glycerol conversions that decayed signif-
conversion with modified/supported heteropoly acids can yield icantly with time. In the first 0.5 h, the glycerol conversions for
excellent activities but the reaction is sometimes challenging due H-BEA, La/HBEA-1 and La/H-BEA-2 were 98%, 95% and 95%, re-
to low catalyst lifetime. This normally occur from the blockage of spectively. However, after 7 h these values reduced to 29.3%, 51.05%
active surface sites by derivatives of oligomerization or other high and 20%, respectively. The acrolein yields also followed a consistent
boiling unsaturated species [112,136,146]. However, recent studies pattern. The respective initial yields were 76.4%, 82.9% and 31.35%
have indicated that supporting the heteropoly acids with shape but decreased to 13.4%, 33.4% and 4.02% after 7 h, respectively. Ac-
selective materials such as zeolites, nano-ZrO2 and silica-alumina cording to these results, La/H-BEA-1 catalyst with moderate acid-
can reduce the chances of active sites blockage and improve the ity and B/L ratio produced the best yield of acrolein and stabil-
overall catalytic stability and activity properties [147-149]. ity properties. Relying on previous studies [76,77,178–180], the au-
thors suggested a mechanism involving the participation of both
4.3. Zeolite catalysts Brønsted and Lewis acid sites, especially due to the nature of the
detected reaction species such as acetol, propanal, carbon deposits
Zeolite catalysts represent a category of shape-selective materi- and others like acetaldehyde and formaldehyde. Acrolein and ace-
als with diverse applications in both biomass and fossil fuels up- tol were believed to be produced via different routes, the aldehy-
grading. The ability to be synthesized via different and improved des followed the decomposition of acrolein. Dehydration involving
methods and the unique structural properties accounted for their the central/secondary –OH group produces 3-hydroxypropanal in-
wide exploitation compared to other catalysts [150-153]. When ap- termediate that subsequently transformed to acrolein. On the con-
propriately designed, zeolites can yield optimal activities even after trary, a path involving terminal/primary –OH group dehydration
cycles of applications. Glycerol upgrading into acrolein over the ze- yields acetol. The higher yield of acrolein with La/H-BEA-1 cata-
olitic catalysts is considered promising due to a number of factors. lyst implied that, the former route was more prominent with this
The process could be achieved at reduced costs, zeolite catalysts catalyst. Catalyst deactivation was considered to be due to coke
are widely available at the commercial scale and numerous stud- formation as observed by some other authors with other catalysts
ies on the reactions involving zeolites have been documented in [133,178,181]. The high susceptibility to deactivation of La/H-BEA-2
the literature [154-158]. Therefore, their properties are understood that was richer in Brønsted acid sites (i.e. B/L = 4.05) indicated that,
to a high degree. Reactions with zeolite catalysts are mainly de- these sites are more prone to cause deactivation than the Lewis
pendent on structure-acidity properties [159-163], which are sub- acid sites.
ject to modifications via changing synthesis routes, varying Si/Al Recently, Decolatti and co-workers [182] studied the role of
ratio, incorporation of metal or oxide modifier or changing the NaOH treatment of H-ZSM-5 zeolite. Following the modification
activation conditions [164-170]. In glycerol upgrading to acrolein, with 0.5 M NaOH, the mesoporosity of the parent H-ZSM-5 ze-
the modifications are normally directed towards achieving high olite increased from 0.254 to 0.325 mL/g whereas the Si/Al ratio
acrolein yields by enhancing acidity properties. Another important decreased from 15.2 to 13.9. On the other hand, the volume of
issue is that, it is not clear as which concentration of the acid sites micropores slightly reduced from 0.136 to 0.130 mL/g. Decreasing

Please cite this article as: A. Galadima, O. Muraza, A review on glycerol valorization to acrolein over solid acid catalysts, Journal of the
Taiwan Institute of Chemical Engineers (2016), http://dx.doi.org/10.1016/j.jtice.2016.07.019
JID: JTICE
ARTICLE IN PRESS [m5G;August 10, 2016;13:11]

10 A. Galadima, O. Muraza / Journal of the Taiwan Institute of Chemical Engineers 000 (2016) 1–16

the zeolite Si/Al ratio is a known factor responsible for increas- 52.13% and 33.04% at the temperatures of 280, 300 and 320 °C,
ing the acidity properties. Thus, following the treatment, the acid- respectively. Detection of propionic acid, formaldehyde and ace-
ity increased from 0.592 to 1.131 mmol/g. Correlation studies re- tone among other products indicated the reaction to possibly in-
vealed features consistent with both Brønsted and Lewis acid sites volved both Brønsted and Lewis acid sites as suggested elsewhere
[183-187]. The B/L ratio for the fresh/parent and NaOH treated H- [171]. However, some authors consider the Brønsted acid sites as
ZSM-5 catalysts following pyridine adsorption at 100 °C were 1.83 the most favorable for acrolein yield [178,194]. Further studies con-
and 2.21, respectively. Therefore, the treated catalyst was denser ducted on the effect of acidity showed the catalytic activity to de-
in Brønsted acid sites. Glycerol upgrading was conducted at 275 °C crease drastically when the Si/Al ratio was increased (see Fig. 5).
using 20 wt% in water and 3 h−1 space velocity for a period of At 320 °C, the conversion decreased from 99.75% for the Si/Al ratio
3 h. Under constant conditions, the treated catalyst was more ac- of 30 to 45.48 and 51.41% for the Si/Al ratios of 50 and 80, respec-
tive. The parent H-ZSM-5 produced glycerol conversion of 45.1% tively. The acrolein yield also reduced from 49.92% to 11.29% and
that reduced to 24.3% in 3 h, whereas for the treated catalyst the 8.24%, respectively. The trend of coke deposition (i.e. 24.61% and
conversion reduced from 63.5% to 31.5%. On the other hand, the 12.87% for the most and least acidic catalysts, respectively) indi-
acrolein selectivity of the parent H-ZSM-5 decreased from 32.3% to cated that, acidity play a great role in enhancing coke formation.
19.0%, whereas 55.2% to 26.1% for the NaOH treated catalyst. A sim- The role of zeolitic channel structure was studied by Gu et al.
ilar behavior was observed when the reactions were repeated at a [195], who employed H-ZSM-5, H-BEA and H-Y zeolites for glycerol
space velocity of 0.75 h−1 . The initial conversions for the parent activation. When the reactions were conducted at 320 °C using 1 g
and treated catalysts were 61.2% and 89.5% and reduced to 13.8% of catalyst, 20 wt% of glycerol in water and 873 h−1 space veloc-
and 77.6% after 5 h, respectively. The respective acrolein selectivi- ity for a period of 1 h, the catalytic activity followed the order H-
ties were 30.6% and 72.1%, but reduced to 4.5% and 58.6%. An anal- Y < H-BEA < H-ZSM-5, whereas the susceptibility to coke forma-
ysis of coke depositions revealed an increased in coke formation tion followed the reverse order. Coke deposits were produced due
with increasing acidity. Thus, the treated catalyst produced more to increasing catalyst acidity. These species blocked the pores and
carbonaceous deposits due to oligomerization than the parent cat- cover the active acid sites with a net decreased in catalytic activ-
alyst. At the space velocities of 0.75 and 3 h−1 , the coke deposits ity. The most active H-ZSM-5 catalyst produced glycerol conversion
for the parent H-ZSM-5 were 12.85% and 13.0% but increased to of 71.8% compared to 46.6 and 69.5% obtained with the H-Y and
14.76% and 17.63% for the treated catalyst, respectively. However, H-BEA catalysts, respectively. Similarly, the respective acrolein se-
the higher activities with the treated catalyst entailed that, coke lectivities were 42.2%, 35.5% and 38.9% for the H-ZSM-5, H-Y and
species were deposited in the inner surface, limiting pore lock- H-BEA catalysts (see Fig. 6). Promising increased in glycerol con-
age due to the enhanced mesoporosity. Glycerol upgrading with version to 93.4% and acrolein selectivity to 52.4% were observed
H-ZSM-5 catalysts have been considered more effective when the when the H-ZSM-5 catalyst was designed in the nanoscale (i.e. 50–
catalysts are designed to the nanoscale with enhanced Brønsted 500 nm). According to the various results, channeling and associ-
acidity [188]. ated diffusion resistance play a key role during the reaction as also
The role of modifying acidity by changing the Si/Al ratio was observed previously with zeolites [196,197]. Bulk H-ZSM-5, H-Y and
also demonstrated by Carrico et al. [189]. H-MCM-22 zeolites with H-BEA exhibited higher diffusion resistance than the nano-H-ZSM-
Si/Al ratios of 30, 50 and 80 were developed through hydrother- 5 catalyst. The H-Y catalyst with the highest diffusion resistance
mal synthesis by adjusting NaOH concentrations, a procedure pre- yields the lowest catalytic activity. Channels of the 10-MR HZSM-
viously proved effective by other authors [190]. The catalyst rela- 5 system were basically both straight of dimension 0.51 × 0.55 nm
tive crystallinity decreased from 100% for the Si/Al ratio of 30 to and intersecting sinusoidal with 0.53 × 0.56 nm. On the other hand,
93 and 81% for the Si/Al ratios of 50 and 80, respectively. On a H-BEA and H-Y are 12-MR systems possessing three-dimensional
similar trend, the volume of mesopores decreased from 0.285 to pore system. H-Y is FAU in nature, exhibiting pores running per-
0.113 and 0.074 mL/g, respectively. Increasing the Si/Al, as usual pendicularly. Large cavity and pore diameter of the H-Y could be
decreased the catalyst acidity. The acidity values were 1.19, 0.76 up to 1.3 and 0.74 nm, respectively. Therefore, the authors [195] ar-
and 0.70 mmol/g for the Si/Al ratios of 30, 50 and 80, respectively. gued that, low channel parameters (i.e. complexity, diameter and
Correlating with other works [191-193], all the catalysts possessed particle size) are necessary factors for enhancing acrolein yields
both strong and weak acid sites whose concentrations were de- with zeolite catalysts. However, a contrary observation was re-
pendent on the catalyst nature. The concentrations of strong acid cently reported following a study by de Oliveira and co-workers
sites for the Si/Al ratios of 30, 50 and 80 were 0.58, 0.39 and [198]. Their results showed an activity trend in the order H-SBA-15
0.30 mmol/g, respectively. Similarly, the concentrations of weak < H-MOR < H-BEA < H-Y. The H-Y zeolite produced the highest
acid sites were 0.61, 0.37 and 0.08 mmol/g, respectively. The trend activity of 89% glycerol conversion and 99.5% acrolein selectivity
indicated decreasing both strong and weak acid sites with increas- due to its large pore structure and higher acidity. The least acidic
ing Si/Al ratio. The most acidic H-MCM-22 (i.e. Si/Al = 30) was em- H-SBA-15 catalyst yielded 40.6% conversion and 84% acrolein selec-
ployed as a model catalyst for glycerol upgrading at temperatures tivity. The reactions were conducted at 250°C using 0.1 g of catalyst
in the range of 280–320 °C, using 36.6 wt% of glycerol in water and and 20 wt% of glycerol in water for 10 h. The H-BEA catalyst yields
0.1 g of catalyst for a period of 12 h. the catalytic activity increased only 18.7% selectivity to acrolein at a conversion closer to that of
substantially with temperature, although a linear reduction was the H-Y zeolite.
observed with time at all temperatures. The initial glycerol con- The effect of water content was demonstrated by Kim and
versions were 57.14%, 69.76% and 99.75% at 280, 300 and 320 °C, co-workers [199], using H-ZSM-5 catalyst having Si/Al of 30.
respectively. These values, however reduced to 33.45%, 8.69% and Glycerol upgrading was performed at 315 °C using 0.3 g of catalyst
20.20% after 12 h, respectively. The acrolein yields were 12.13%, and water concentrations of 15.7 and 51.9 mol%. Increasing the
26.04% and 49.92%, but reduced to 2.82%, 2.48% and 4.30%, re- concentration of water (i.e. decreasing glycerol concentration) fa-
spectively. Coke deposition also increased with increase in reac- vors catalytic activity. Glycerol conversion was 45.8% at the water
tion temperature. At 280 °C, the amount of coke was 18.78% but concentration of 15.7 mol% but increased to 51.1% when the water
increased to 22.08% and 24.61% at 300 and 320 °C, respectively. concentration was raised to 51.9 mol% within 2 h. However, these
The high coke deposits could be attributed to the strong acid- values decreased to 26.3% and 25.6% for the 15.7% and 51.9 mol% of
ity properties of this catalyst. Among the side reaction products, water after 12 h, respectively. Yield of acrolein was also favored by
CH4 and COx (x = 1 or 2) dominated with selectivities of 66.56%, higher water concentrations. With 15.7 mol % water, the yield was

Please cite this article as: A. Galadima, O. Muraza, A review on glycerol valorization to acrolein over solid acid catalysts, Journal of the
Taiwan Institute of Chemical Engineers (2016), http://dx.doi.org/10.1016/j.jtice.2016.07.019
JID: JTICE
ARTICLE IN PRESS [m5G;August 10, 2016;13:11]

A. Galadima, O. Muraza / Journal of the Taiwan Institute of Chemical Engineers 000 (2016) 1–16 11

Fig. 5. Effect of H-MCM-22 acidity on glycerol conversion and acrolien yield. Reaction conditions: 320 °C, 36.6 wt% of glycerol in water, 0.1 g for 12 h. Data source: Ref. [189].

Fig. 6. Effect of zeolite structure on glycerol conversion and acrolien selectivity. Reaction conditions: 320 °C, 1 g of catalyst, 20 wt% of glycerol in water and 873 h−1 . Data
source: Ref. [195].

3.3% but increased to 12.5% when the water content was raised to must carefully be selected otherwise the reaction would produce
51.9 mol%. However, the nature/type of the other reaction products aromatics and their alkylated derivatives as the dominant reaction
was the same at all water concentrations. Propionic acid, acetone products [202]. Recent studies involving the zeolite catalysts
and hydroxyacetone were all detected to very low selectivity (max. have shown their textural properties and the nature of Brønsted
10%). Other unidentified reaction products were produced to very acidity (i.e. strong or weak) coupled with the reaction conditions
high selectivity (i.e. up to 80%). Irrespective of the water concentra- as very influencial in achieving enhanced catalytic activity and
tion, the carbon deposits was constant at ∼ 10%. Although different stability [203-205]. Zeolites with more open pore-structure are
strategies including the introduction of oxygen into the reaction good candidates for limiting coke deposition due to elimination of
feed, changing catalyst acidity and modifying reactor configuration diffusional problems. On the other hand, the moderate to strong
have been considered useful [194, 199, 200], glycerol upgrading Brønsted acid sites facilitate the dehydration of glycerol with a
with zeolites is still problematic due to the zeolites susceptibility net consequence om improved acrolein yield, especially when the
to carbonaceous depositions associated with textural and acidity reaction is performed at temperatures between 250 and 350 °C.
properties. This creates difficulty during the reaction. Therefore,
developing new strategies for ensuring long term catalyst stability 4.4. Prospects of silicoalumino phosphate (SAPO) catalysts
is a very critical factor. Thermal stability is very important both for
the reaction and catalyst recycling purposes. Another important Silicoaluminophosphates are another class of molecular sieves
issue to be adequately explored is the influence of water concen- that have recently attracted industrial attention, especially in the
tration on the zeolite structure, acidity and stability properties. areas of catalysis. Like the zeolites, their structural and acidity
Although the participation of both Brønsted and Lewis acids have properties could be modified to model them for specific reactions.
been demonstrated, it is still not clear what actual positive role They have been used as good candidates for processes such as
the Lewis acid sites play in enhancing acrolein yield and limiting ion exchange, active or support components for catalysts and dur-
catalyst deactivation by coking. Glycerol from biodiesel produc- ing purification of molecular derivatives [206]. Many SAPOs have
tion may contain trace quantities of undesirable impurities like been synthesized including those with chabazite, levynite and fauj-
catalyst particles, water and short chain alcohol (like methanol). asite structures. The ability of these systems to be designed having
Their effect (s) on catalyst performance towards acrolein yield variable cage sizes, windows and acidity properties permits their
needs to be adequately evaluated. Recently, da Silva and Mota exploration for many structure-acidity dependent reactions. They
[201] have reported these impurities to significantly hindered (i.e. have shown good activities in methanol to hydrocarbon reactions,
by up to 50%) the reaction of glycerol with acetone over H-BEA including methanol to gasoline (MTG) and methanol to olefins
zeolite and Amberlyst-15 catalysts. Another issue that should be (MTO) reactions due to their hierarchical nature with reduced-
considered is that the reaction conditions and zeolite topology diffusion limitations [207]. High conversions (> 90%) of methanol

Please cite this article as: A. Galadima, O. Muraza, A review on glycerol valorization to acrolein over solid acid catalysts, Journal of the
Taiwan Institute of Chemical Engineers (2016), http://dx.doi.org/10.1016/j.jtice.2016.07.019
JID: JTICE
ARTICLE IN PRESS [m5G;August 10, 2016;13:11]

12 A. Galadima, O. Muraza / Journal of the Taiwan Institute of Chemical Engineers 000 (2016) 1–16

and olefins selectivity (up to 90%) have been reported with pure or and 5% after 10 h, respectively. On the other hand, initial acrolein
modified SAPOs like SAPO-34 [207–209], metals promoted SAPO- selectivities were 60% and 75% but reduced to 20% and 10%, respec-
34 [210,211] and other SAPO-n (n = 11,18,35 etc) [212,213], at tem- tively. According to these results, SAPO-11 demonstrated higher
peratures closer to those being employed for glycerol upgrading. stability under constant conditions. This can be observed from the
They are also associated with longer catalyst lifetime when ap- amount of coke (i.e. 5.4 and 6.6 wt%) deposited for the two respec-
propriately modified [214-216]. Similar properties have been doc- tive catalysts. Among other reaction products, acetol was produced
umented for the MTG reaction with these catalysts. In a trend to the highest selectivity of 6.8% to 10%. Phenol was detected with
similar to the glycerol upgrading, the MTG and MTO reactions SAPO-11 but not SAPO-34. Glycidol and acetaldehyde were all de-
are Brønsted acid sites mediated reactions [217]. However, de- tected to very negligible selectivity. The nature of the reaction
spite these unique properties, only few recent studies have been products showed the reaction to proceeds through similar mech-
documented on their activities in glycerol upgrading to acrolein anism reported by Lourenco and co-workers [221], but with sub-
[218–220]. sequent side reactions involving acrolein, 3-hydroxypropanal and
A study by Lourenco and co-workers [221], employed SAPO-n acetol.
(n = 11, 34 and 40) and demonstrated that acidity and porous struc- The few glycerol to acrolein studies conducted so far with
ture played key roles during the reaction. The amounts of Brønsted SAPOs indicated the reaction mechanism and role of reaction
and Lewis acid sites for the catalysts ranged from 0.046 to 0.28 conditions to follow similar patterns with other solid acid cata-
and 0.049 to 0.056 mmol/g, respectively, and increased in the or- lysts discussed earlier. However, an appropriate choice of SAPO-
der SAPO-34 < SAPO-11 < SAPO-40. Two unique reaction path- structure and acidity properties can enhance glycerol conversion,
ways were observed with these catalysts in a manner similar to acrolein selectivity and catalyst lifetime. There are vital issues that
other solid acid catalysts. The desired reaction pathway proceeds should be adequately explored to fully establish the potentials of
with sequential dehydration of glycerol to 3-hydroxypropanal and the SAPO catalysts. These include studying the effects of modifica-
acrolein whereas the alternative pathway involved single dehydra- tions with metals (e.g. like Ce and La), incorporation of support like
tion step to acetol. Initial model reactions conducted with SAPO-40 titania, zirconia, alumina and silica and exploring other numerous
using 10 wt% of glycerol in water and space velocity of 0.85 h−1 for SAPO structures. The role of catalyst preparation, crystallite sizes
120 h, indicated the reaction to be temperature dependent. Increas- and pre-treatment conditions and adequate characterization of the
ing the reaction temperature from 320 to 380 °C increased glyc- coke deposits can provide further insights into their prospects as
erol conversion from 89% to 100%. The conversions were mainly glycerol upgrading catalysts.
stable for about 60 h but decreased by 50% after 120 h of reac-
tion. On the other hand, the catalyst produced stable acrolein se- 5. Conclusions
lectivities of 75.8%, 73% and 74.5% at 320, 350 and 380 °C after the
120 h. Combined selectivity to acetol and formadehyde ranged be- Glycerol provides a green (i.e. environmentally sustainable)
tween 12% and 20% at all the temperatures. However, up to 20% option for the production of acrolein to very good yields. Protonic
of coke was deposited when the temperature reached 380 °C for catalysts such as mixed and modified acidic oxides, heteropoly
120 h. The deactivation can however be considered milder com- acids and zeolites have been recently studied as catalysts for
pared to > 20% coke depositions reported other solid catalysts dis- upgrading glycerol to acrolein. Although there is a unanimous
cussed above, at lower reaction temperatures and times. When the literature agreement on the role Brønsted acid sites (i.e. as main
reactions were repeated at 350 °C using 30 wt% of glycerol in wa- active sites), certain issues related to the mechanism are not yet
ter for 2.5 h, space velocity played an important role. At the space resolved. These include the densities of these sites that yield
velocities of 0.86 and 2.68 h−1 , the conversion was stable at 100% optimal activity (i.e. > 90% acrolein yield) with specific catalyst,
but reduced to 79% when the space velocity was further raised which –OH group (i.e. primary or secondary) in the glycerol is first
to 4.69 h−1 . Among these conditions, space velocity of 2.68 h−1 dehydrated and associated role on acrolein versus side product
yields the optimal acrolein selectivity of 70% compared to 63.7% selectivity and the actual role of Lewis acid sites. Others include
and 68.1% obtained with space velocities of 0.86 and 4.69 h−1 , details on the mechanism of formation of other side reaction prod-
respectively. Comparative study conducted with other SAPO cat- ucts like phenol and the full identification of unknown compounds
alysts (i.e. SAPO-11 and SAPO-34) at 350 °C, 0.85 h−1 and 30 wt% usually produced when the reaction is conducted with catalysts of
of glycerol in water, showed SAPO-11 and SAPO-40 as the most low Brønsted acidity. Oxide catalysts have demonstrated good ac-
active catalysts, producing 100% glycerol conversion compared to tivity and acrolein selectivity properties, but further studies on the
90% obtained with the SAPO-34. Their acrolein selectivities were actual role of textural properties and the influence of air or nitro-
appreciably similar (> 70%) but the reduction in glycerol conver- gen during their activation would be very critical. SO4 2− modified
sion was severe with SAPO-11 and severest with SAPO-34. The oxides like ZrO2 have so far demonstrated low activity due to loss
results of the study revealed that, although all the catalysts pro- of the SO4 2− species (i.e. the main Brønsted acid sites generating
duced similar types of reaction products and uniquely high con- species) during the reaction. Therefore, effort to develop appropri-
versions and selectivities, structure and acidity played vital roles. ate stabilizing agents would be crucial in promoting the activity
SAPO-11 and SAPO-40 being more acidic produced the best glyc- with these catalysts. Heteropoly acids are associated with suffi-
erol upgrading activities. The SAPO-34 (8-MR, cage-like structure) cient acidity and textural properties to shift the reaction to high
with small pore properties deactivated more rapidly than SAPO- acrolein yields. However, these catalysts demonstrated poor cata-
11 (10-MR, channel-structure) with intermediate pore properties. lyst lifetime in a number of cases due to their ability to catalyze
SAPO-40 (12-MR, interconnected-structure) exhibited the highest oligomerization and the formation of high boiling unsaturated
stability properties for 120 h. Therefore this catalyst could further compounds, with strong potentials for coke depositions. Zeolite
be evaluated in comparison to other solid acid catalysts like zeo- catalysts demonstrated unique activities due to the possibility
lites and heteropoly acids. of modifying the structural/acidity properties and their well-
Suprun et al. [222] conducted a study with SAPO-11 and SAPO- documented records in Brønsted acid sites mediated reactions.
34 having 0.5 and 1.335 mmol/g total acid sites, respectively. When However, the positive role of the Lewis acid sites during the reac-
glycerol conversions were conducted at 280 °C using 5 wt% of glyc- tion, effect of impurities on catalyst performance and the possible
erol in water and 43 h−1 for 10 h, the initial glycerol conversions ways of inhibiting coke deposition (i.e. enhancing catalyst lifetime)
for SAPO-11 and SAPO-34 were 83% and 70% but decreased to 20% need to be fully addressed. Another issue with these catalysts

Please cite this article as: A. Galadima, O. Muraza, A review on glycerol valorization to acrolein over solid acid catalysts, Journal of the
Taiwan Institute of Chemical Engineers (2016), http://dx.doi.org/10.1016/j.jtice.2016.07.019
JID: JTICE
ARTICLE IN PRESS [m5G;August 10, 2016;13:11]

A. Galadima, O. Muraza / Journal of the Taiwan Institute of Chemical Engineers 000 (2016) 1–16 13

is the lack of unanimous details on which zeolite structure and [23] de Sousa FF, Oliveira AC, Filho JM, Pinheiro GS, Giotto M, Barros NA,
acidity combination provide the best activities. Very limited stud- et al. Metal oxides nanoparticles from complexes on SBA-15 for glycerol con-
version. Chem Eng J 2013;228:442–8.
ies are available on SAPO catalysts. Although these materials have [24] Feng Y, Yin H, Wang A, Shen L, Yu L, Jiang T. Gas phase hydrogenolysis of
so far demonstrated improved performance, there are critical is- glycerol catalyzed by Cu/ZnO/MOx (MOx = Al2 O3 , TiO2 , and ZrO2 ) catalysts.
sues to be explored. These include the effects of modification with Chem Eng J 2011;168:403–12.
[25] Zakaria ZY, Linnekoski J, Amin NAS. Catalyst screening for conversion of glyc-
metals or catalyst supports, role of catalyst preparation conditions, erol to light olefins. Chem Eng J 2012;207–208:803–13.
SAPO structure and acid site densities on catalytic activity. [26] Pérez-Barrado E, Pujol MC, Aguiló M, Llorca J, Cesteros Y, Díaz F, et al. Influ-
ence of acid–base properties of calcined MgAl and CaAl layered double hy-
droxides on the catalytic glycerol etherification to short-chain polyglycerols.
Acknowledgment Chem Eng J 2015;264:547–56.
[27] Bart JCJ, Palmeri N, Cavallaro S, 13 - Valorisation of the glycerol by-product
from biodiesel production. In: J.C.J. Bart, N. Palmeri, S. Cavallaro (Eds.)
The authors would like to acknowledge the funding provided by Biodiesel Science and Technology, Woodhead Publishing, 2010, pp. 571-624.
King Abdulaziz City for Science and Technology (KACST) through [28] Zakaria ZY, Amin NAS, Linnekoski J. A perspective on catalytic conversion of
glycerol to olefins. Biomass Bioenergy 2013;55:370–85.
the Science & Technology Unit in Center of Research Excellence [29] Muraza O. Maximizing diesel production through oligomerization: a land-
in Nanotechnology at King Fahd University of Petroleum & Min- mark opportunity for zeolite research. Ind Eng Chem Res 2014;54:781–9.
erals (KFUPM) for supporting this work through project No. 13- [30] Markočič E, Kramberger B, van Bennekom JG, Jan Heeres H, Vos J, Knez Ž.
Glycerol reforming in supercritical water; a short review. Renew Sustain En-
NAN1702-04 as part of the National Science, Technology and In- ergy Rev 2013;23:40–8.
novation Plan. [31] Teng WK, Ngoh GC, Yusoff R, Aroua MK. A review on the performance of
glycerol carbonate production via catalytic transesterification: effects of in-
fluencing parameters. Energy Convers Manag 2014;88:484–97.
References [32] Waghmare GV, Vetal MD, Rathod VK. Ultrasound assisted enzyme catalyzed
synthesis of glycerol carbonate from glycerol and dimethyl carbonate. Ultra-
[1] Ming T, de_Richter R, Liu W, Caillol S. Fighting global warming by climate son Sonochem 2015;22:311–16.
engineering: is the Earth radiation management and the solar radiation man- [33] Omata K, Izumi S, Murayama T, Ueda W. Hydrothermal synthesis of W–Nb
agement any option for fighting climate change? Renew Sustain Energy Rev complex metal oxides and their application to catalytic dehydration of glyc-
2014;31:792–834. erol to acrolein. Catal Today 2013;201:7–11.
[2] Bruun S, Jensen LS, Khanh Vu VT, Sommer S. Small-scale household biogas [34] Tomitori H, Nakamura M, Sakamoto A, Terui Y, Yoshida M, Igarashi K,
digesters: an option for global warming mitigation or a potential climate et al. Augmented glutathione synthesis decreases acrolein toxicity. Biochem
bomb? Renew Sustain Energy Rev 2014;33:736–41. Biophys Res Commun 2012;418:110–15.
[3] Bolaji BO, Huan Z. Ozone depletion and global warming: case for the use of [35] Talebian-Kiakalaieh A, Amin NAS, Hezaveh H. Glycerol for renewable
natural refrigerant – a review. Renew Sustain Energy Rev 2013;18:49–54. acrolein production by catalytic dehydration. Renewable Sustain Energy Rev
[4] Lee ZH, Sethupathi S, Lee KT, Bhatia S, Mohamed AR. An overview on global 2014;40:28–59.
warming in Southeast Asia: CO2 emission status, efforts done, and barriers. [36] Deaton WI, Glycerol from allyl alcohol, in, Google Patents, 1963.
Renew Sustain Energy Rev 2013;28:71–81. [37] Wróblewska A, Fajdek A. Epoxidation of allyl alcohol to glycidol over the mi-
[5] Florides GA, Christodoulides P, Messaritis V. Reviewing the effect of CO2 and croporous TS-1 catalyst. J Hazard Mater 2010;179:258–65.
the sun on global climate. Renew Sustain Energy Rev 2013;26:639–51. [38] García-Serna J, Pérez-Barrigón L, Cocero MJ. New trends for design to-
[6] Sorate KA, Bhale PV. Biodiesel properties and automotive system compatibil- wards sustainability in chemical engineering: green engineering. Chem Eng
ity issues. Renew Sustain Energy Rev 2015;41:777–98. J 2007;133:7–30.
[7] Guldhe A, Singh B, Mutanda T, Permaul K, Bux F. Advances in synthesis of [39] Anastas PT, Zimmerman JB. Peer reviewed: design through the 12 principles
biodiesel via enzyme catalysis: novel and sustainable approaches. Renew Sus- of green engineering. Environ Sci Technol 2003;37:94–101.
tain Energy Rev 2015;41:1447–64. [40] Global Group and ADM team up to develop a catalyst to produce glycerol
[8] Takase M, Zhao T, Zhang M, Chen Y, Liu H, Yang L, et al. An expatiate review from carbohydrates. Focus Catal 2013;2013:2.
of neem, jatropha, rubber and karanja as multipurpose non-edible biodiesel [41] Tomasik P, Horton D. Chapter 2 - Enzymatic conversions of starch. In:
resources and comparison of their fuel, engine and emission properties. Re- Derek H, editor. Advances in carbohydrate chemistry and biochemistry, 68.
new Sustain Energy Rev 2015;43:495–520. Academic Press; 2012. p. 59–436.
[9] DÁgosto MdA, Vieira da Silva MA, de Oliveira CM, Franca LS, da Costa [42] Bouaid A, El boulifi N, Hahati K, Martinez M, Aracil J. Biodiesel produc-
Marques LG, Soares Murta AL, et al. Evaluating the potential of the use tion from biobutanol. Improvement of cold flow properties. Chem Eng J
of biodiesel for power generation in Brazil. Renew Sustain Energy Rev 2014;238:234–41.
2015;43:807–17. [43] Liu Y, Tan H, Zhang X, Yan Y, Hameed BH. Effect of monohydric alco-
[10] Bharathiraja B, Chakravarthy M, Kumar RR, Yuvaraj D, Jayamuthunagai J, Ku- hols on enzymatic transesterification for biodiesel production. Chem Eng J
mar RP, et al. Biodiesel production using chemical and biological methods 2010;157:223–9.
– A review of process, catalyst, acyl acceptor, source and process variables. [44] Gu L, Huang W, Tang S, Tian S, Zhang X. A novel deep eutectic solvent
Renew Sustain Energy Rev 2014;38:368–82. for biodiesel preparation using a homogeneous base catalyst. Chem Eng J
[11] Yaakob Z, Narayanan BN, Padikkaparambil S, S Unni K, M Akbar P. A re- 2015;259:647–52.
view on the oxidation stability of biodiesel. Renew Sustain Energy Rev [45] Lam MK, Lee KT. Immobilization as a feasible method to simplify the sep-
2014;35:136–53. aration of microalgae from water for biodiesel production. Chem Eng J
[12] NBB. Production statistics, biodiesel. U.S.A.: National Biodiesel Board; 2014 2012;191:263–8.
http://www.biodiesel.org/production/production-statistics. [46] Sun P, Sun J, Yao J, Zhang L, Xu N. Continuous production of biodiesel from
[13] Hervé G, Yves D, Agneta F. Biofuels and world agricultural markets: outlook high acid value oils in microstructured reactor by acid-catalyzed reactions.
for 2020 and 2050. INTECH Open Access Publisher; 2011. Chem Eng J 2010;162:364–70.
[14] Galadima A, Muraza O. Biodiesel production from algae by using heteroge- [47] Medeiros MA, Leite CMM, Lago RM. Use of glycerol by-product of biodiesel to
neous catalysts: a critical review. Energy 2014;78:72–83. produce an efficient dust suppressant. Chem Eng J 2012;180:364–9.
[15] Quispe CAG, Coronado CJR, Carvalho JA Jr. Glycerol: production, consumption, [48] Berrios M, Skelton RL. Comparison of purification methods for biodiesel.
prices, characterization and new trends in combustion. Renew Sustain Energy Chem Eng J 2008;144:459–65.
Rev 2013;27:475–93. [49] Xie W, Fan M. Biodiesel production by transesterification using tetraalkylam-
[16] Michael Malveda MB. Chiyo funada and jiangli liu glycerin. USA: IHS; 2012 monium hydroxides immobilized onto SBA-15 as a solid catalyst. Chem Eng J
http://chemical.ihs.com/CEH/Public/Reports/662.50 0 0. 2014;239:60–7.
[17] Tapah BF, Santos RCD, Leeke GA. Processing of glycerol under sub and super- [50] Teixeira APC, Santos EM, Vieira AFP, Lago RM. Use of chrysotile to produce
critical water conditions. Renew Energy 2014;62:353–61. highly dispersed K-doped MgO catalyst for biodiesel synthesis. Chem Eng J
[18] Markočič E, Kramberger B, Van Bennekom JG, Jan Heeres H, Vos J, Knez Ž. 2013;232:104–10.
Glycerol reforming in supercritical water: a short review. Renew Sustain En- [51] Taufiq-Yap YH, Lee HV, Yunus R, Juan JC. Transesterification of non-edible Jat-
ergy Rev 2013;23:40–8. ropha curcas oil to biodiesel using binary Ca–Mg mixed oxide catalyst: effect
[19] Zakaria ZY, Amin NAS, Linnekoski J. A perspective on catalytic conversion of of stoichiometric composition. Chem Eng J 2011;178:342–7.
glycerol to olefins. Biomass Bioenergy 2013;55:370–85. [52] Boey P-L, Maniam GP, Hamid SA. Performance of calcium oxide as a heteroge-
[20] González MD, Cesteros Y, Salagre P. Establishing the role of Brønsted acidity neous catalyst in biodiesel production: a review. Chem Eng J 2011;168:15–22.
and porosity for the catalytic etherification of glycerol with tert-butanol by [53] Katada N, Hatanaka T, Ota M, Yamada K, Okumura K, Niwa M.
modifying zeolites. Appl Catal, A 2013;450:178–88. Biodiesel production using heteropoly acid-derived solid acid catalyst
[21] Serafim H, Fonseca IM, Ramos AM, Vital J, Castanheiro JE. Valorization of glyc- H4PNbW11O40/WO3 –Nb2 O5 . Appl Catal, A 2009;363:164–8.
erol into fuel additives over zeolites as catalysts. Chem Eng J 2011;178:291–6. [54] Sree R, Kuriakose S. Alkali salts of heteropoly tungstates: efficient catalysts
[22] Al-Lal AM, García-González JE, Llamas A, Monjas A, Canoira L. A new route to for the synthesis of biodiesel from edible and non-edible oils. J Energy Chem
synthesize tert-butyl ethers of bioglycerol. Fuel 2012;93:632–7. 2015;24:87–92.

Please cite this article as: A. Galadima, O. Muraza, A review on glycerol valorization to acrolein over solid acid catalysts, Journal of the
Taiwan Institute of Chemical Engineers (2016), http://dx.doi.org/10.1016/j.jtice.2016.07.019
JID: JTICE
ARTICLE IN PRESS [m5G;August 10, 2016;13:11]

14 A. Galadima, O. Muraza / Journal of the Taiwan Institute of Chemical Engineers 000 (2016) 1–16

[55] Talebian-Kiakalaieh A, Amin NAS, Zarei A, Noshadi I. Transesterification of [87] Feng X, Yao Y, Su Q, Zhao L, Jiang W, Ji W, et al. Vanadium pyrophosphate
waste cooking oil by heteropoly acid (HPA) catalyst: optimization and kinetic oxides: The role of preparation chemistry in determining renewable acrolein
model. Appl Energy 2013;102:283–92. production from glycerol dehydration. Appl Catal, B 2015;164:31–9.
[56] Singh D, Ganesh A, Mahajani S. Heterogeneous catalysis for biodiesel [88] Chheda JN, Huber GW, Dumesic JA. Liquid-phase catalytic processing of
synthesis and valorization of glycerol. Clean Technol Environ Policy biomass-derived oxygenated hydrocarbons to fuels and chemicals. Angew
2014;17:1103–10. Chem - Int Ed 2007;46:7164–83.
[57] Hassani M, Najafpour GD, Mohammadi M, Rabiee M. Preparation, character- [89] Feng X, Sun B, Yao Y, Su Q, Ji W, Au CT. Renewable production of acrylic acid
ization and application of zeolite-based catalyst for production of biodiesel and its derivative: new insights into the aldol condensation route over the
from waste cooking oil. J Sci Ind Res 2014;73:129–33. vanadium phosphorus oxides. J Catal 2014;314:132–41.
[58] Wang Y-Y, Wang H-H, Chuang T-L, Chen B-H, Lee D-J. Biodiesel produced [90] Wang F, Dubois J-L, Ueda W. Catalytic performance of vanadium pyrophos-
from catalyzed transesterification of triglycerides using ion-exchanged zeolite phate oxides (VPO) in the oxidative dehydration of glycerol. Appl Catal, A
beta and MCM-22. Energy Proc 2014;61:933–6. 2010;376:25–32.
[59] Wang Y-Y, Lee D-J, Chen B-H. Low-Al zeolite beta as a heterogeneous catalyst [91] Liu Q, Zhang Z, Du Y, Li J, Yang X. Rare earth pyrophosphates: effective cata-
in biodiesel production from microwave-assisted transesterification of triglyc- lysts for the production of acrolein from vapor-phase dehydration of glycerol.
erides. Energy Proc 2014;61:918–21. Catal Lett 2009;127:419–28.
[60] Feyzi M, Khajavi G. Investigation of biodiesel production using modified [92] Wang F, Dubois J-L, Ueda W. Catalytic dehydration of glycerol over vana-
strontium nanocatalysts supported on the ZSM-5 zeolite. Ind Crops Prod dium phosphate oxides in the presence of molecular oxygen. J Catal
2014;58:298–304. 2009;268:260–7.
[61] Borges LD, Moura NN, Costa AA, Braga PRS, Dias JA, Dias SCL, et al. Inves- [93] Cecilia JA, García-Sancho C, Mérida-Robles JM, Santamaría-González J,
tigation of biodiesel production by HUSY and Ce/HUSY zeolites: influence of Moreno-Tost R, Maireles-Torres P. V and V–P containing Zr-SBA-15 catalysts
structural and acidity parameters. Appl Catal, A 2013;450:114–19. for dehydration of glycerol to acrolein 2015;254:43–52.
[62] Stevens JF, Maier CS. Acrolein: sources, metabolism, and biomolecular in- [94] Stošić D, Bennici S, Sirotin S, Calais C, Couturier J-L, Dubois J-L, et al. Glycerol
teractions relevant to human health and disease. Mol Nutr Food Res dehydration over calcium phosphate catalysts: effect of acidic–basic features
2008;52:7–25. on catalytic performance. Appl Catal, A 2012;447–448:124–34.
[63] Yadav GD, Sharma RV, Katole SO. Selective Dehydration of glycerol to [95] Bagheri S, Julkapli NM, Yehye WA. Catalytic conversion of biodiesel de-
acrolein: development of efficient and robust solid acid catalyst MUICaT-5. rived raw glycerol to value added products. Renew Sustain Energy Rev
Ind Eng Chem Res 2013;52:10133–44. 2015;41:113–27.
[64] Carrazán SRG, Martín C, Mateos R, Rives V. Influence of the active phase [96] Chai S-H, Yan B, Tao L-Z, Liang Y, Xu B-Q. Sustainable production of acrolein:
structure Bi-Mo-Ti-O in the selective oxidation of propene. Catal Today catalytic gas-phase dehydration of glycerol over dispersed tungsten oxides on
2006;112:121–5. alumina, zirconia and silica. Catal Today 2014;234:215–22.
[65] Zhang X, Wan H, Weng W. Reaction pathways for selective oxida- [97] Kim T, Burrows A, Kiely CJ, Wachs IE. Molecular/electronic structure-sur-
tion of propane to acrolein over Ce-Ag-Mo-P-O catalysts. Appl Catal, A face acidity relationships of model-supported tungsten oxide catalysts. J Catal
2009;353:24–31. 2007;246:370–81.
[66] Zhang X, Wan H-l, Weng W-z, Yi X-d. Selective Oxidation of propane to [98] Scheithauer M, Grasselli RK, Knözinger H. Genesis and structure of WOx /ZrO2
acrolein over Ce-doped Ag–Mo–P–O catalysts: influence of Ce promoter. Catal solid acid catalysts. Langmuir 1998;14:3019–29.
Lett 2003;87:229–34. [99] Barton DG, Soled SL, Iglesia E. Solid acid catalysts based on supported tung-
[67] Liu J, Mohamed F, Sauer J. Selective oxidation of propene by vanadium oxide sten oxides. Top Catal 1998;6:87–99.
monomers supported on silica. J Catal 2014;317:75–82. [100] Wachs IE, Kim T, Ross EI. Catalysis science of the solid acidity of model sup-
[68] Rozanska X, Fortrie R, Sauer J. Size-dependent catalytic activity of supported ported tungsten oxide catalysts. Catal Today 2006;116:162–8.
vanadium oxide species: oxidative dehydrogenation of propane. J Am Chem [101] Massa M, Andersson A, Finocchio E, Busca G. Gas-phase dehydration of glyc-
Soc 2014;136:7751–61. erol to acrolein over Al2 O3 -, SiO2 -, and TiO2 -supported Nb- and W-oxide cat-
[69] Dai GL, Li ZH, Lu J, Wang WN, Fan KN. Deep oxidations in the oxidative de- alysts. J Catal 2013;307:170–84.
hydrogenation reaction of propane over V 2O 5(001): periodic density func- [102] Kim YT, Jung KD, Park ED. Gas-phase dehydration of glycerol over silica-alu-
tional theory study. J Phys Chem C 2012;116:807–17. mina catalysts. Appl Catal, B 2011;107:177–87.
[70] Lo JMH, Premji ZA, Ziegler T, Clark PD. First-principles investigation of selec- [103] Suprun W, Lutecki M, Gläser R, Papp H. Catalytic activity of bifunctional tran-
tive oxidation of propane on clean and sulfided V2O5 (010) surfaces. J Phys sition metal oxide containing phosphated alumina catalysts in the dehydra-
Chem C 2013;117:11258–74. tion of glycerol. J Mol Catal A: Chem 2011;342-343:91–100.
[71] Schuh K, Kleist W, Høj M, Trouillet V, Beato P, Jensen AD, et al. Selective [104] Stošić D, Bennici S, Couturier JL, Dubois JL, Auroux A. Influence of surface
oxidation of propylene to acrolein by hydrothermally synthesized bismuth acid-base properties of zirconia and titania based catalysts on the product
molybdates. Appl Catal, A 2014;482:145–56. selectivity in gas phase dehydration of glycerol. Catal Commun 2012;17:23–8.
[72] Liu C-H, Lai N-C, Lee J-F, Chen C-S, Yang C-M. SBA-15-supported highly dis- [105] Massa M, Andersson A, Finocchio E, Busca G, Lenrick F, Wallenberg LR. Per-
persed copper catalysts: vacuum–thermal preparation and catalytic studies in formance of ZrO2 -supported Nb- and W-oxide in the gas-phase dehydration
propylene partial oxidation to acrolein. J Catal 2014;316:231–9. of glycerol to acrolein. J Catal 2013;297:93–109.
[73] Liu L, Ye XP, Bozell JJ. A comparative review of petroleum-based and [106] Lauriol-Garbay P, Millet JMM, Loridant S, Bellière-Baca V, Rey P. New efficient
bio-based acrolein production. ChemSusChem 2012;5:1162–80. and long-life catalyst for gas-phase glycerol dehydration to acrolein. J Catal
[74] Kongpatpanich K, Nanok T, Boekfa B, Probst M, Limtrakul J. Structures and 2011;280:68–76.
reaction mechanisms of glycerol dehydration over H-ZSM-5 zeolite: a density [107] Martínez MT, Martínez MD, Osácar J, Miranda J. Coal liquefaction, deactiva-
functional theory study. Phys Chem Chem Phys 2011;13:6462–70. tion of coal hydrogenation catalysts. Fuel Process Technol 1988;18:51–8.
[75] Chai S-H, Wang H-P, Liang Y, Xu B-Q. Sustainable production of acrolein: in- [108] Clacens JM, Pouilloux Y, Barrault J. Selective etherification of glycerol to
vestigation of solid acid-base catalysts for gas-phase dehydration of glycerol. polyglycerols over impregnated basic MCM-41 type mesoporous catalysts.
Green Chem 2007;9:1130–6. Appl Catal, A 2002;227:181–90.
[76] Corma A, Huber GW, Sauvanaud L, O’Connor P. Biomass to chemicals: cat- [109] Rao Ginjupalli S, Mugawar S, P Rajan N, Kumar Balla P, Chary Komandur VR.
alytic conversion of glycerol/water mixtures into acrolein, reaction network. J Vapour phase dehydration of glycerol to acrolein over tungstated zirconia cat-
Catal 2008;257:163–71. alysts. Appl Surf Sci 2014;309:153–9.
[77] Tsukuda E, Sato S, Takahashi R, Sodesawa T. Production of acrolein from glyc- [110] Lauriol-Garbey P, Loridant S, Bellière-Baca V, Rey P, Millet J-MM. Gas
erol over silica-supported heteropoly acids. Catal Commun 2007;8:1349–53. phase dehydration of glycerol to acrolein over WO 3/ZrO 2 catalysts: im-
[78] Konaka A, Tago T, Yoshikawa T, Shitara H, Nakasaka Y, Masuda T. Conversion provement of selectivity and stability by doping with SiO2 . Catal Commun
of biodiesel-derived crude glycerol into useful chemicals over a zirconia–iron 2011;16:170–4.
oxide catalyst. Ind Eng Chem Res 2013;52:15509–15. [111] Katryniok B, Paul S, Bellière-Baca V, Rey P, Dumeignil F. Glycerol dehy-
[79] Watanabe M, Iida T, Aizawa Y, Aida TM, Inomata H. Acrolein synthesis from dration to acrolein in the context of new uses of glycerol. Green Chem
glycerol in hot-compressed water. Bioresour Technol 2007;98:1285–90. 2010;12:2079–98.
[80] Ott L, Bicker M, Vogel H. Catalytic dehydration of glycerol in sub- and super- [112] Alhanash A, Kozhevnikova EF, Kozhevnikov IV. Gas-phase dehydration of
critical water: a new chemical process for acrolein production. Green Chem glycerol to acrolein catalysed by caesium heteropoly salt. Appl Catal, A
2006;8:214–20. 2010;378:11–18.
[81] Akizuki M, Oshima Y. Kinetics of glycerol dehydration with WO3 /TiO2 in su- [113] Gan H, Zhao X, Song B, Guo L, Zhang R, Chen C, et al. Gas phase dehydra-
percritical water. Ind Eng Chem Res 2012;51:12253–7. tion of glycerol to acrolein catalyzed by zirconium phosphate. Chin J Catal
[82] Schwenk E GM, Aichner F, assigned to Schering-Kahlbaum AG, 1933 US Patent 2014;35:1148–56.
1916743; (1933). [114] Tao L-Z, Chai S-H, Zuo Y, Zheng W-T, Liang Y, Xu B-Q. Sustainable production
[83] Neher HT, Dietrich A, Klenk H, Girke W. (Degussa), Process for the production of acrolein: acidic binary metal oxide catalysts for gas-phase dehydration of
of acrolein, 1994, DE patent 4238493; (1994). glycerol. Catal Today 2010;158:310–16.
[84] Neher HT, Dietrich A, Klenk H, Girke W. (Degussa)„ Process for the production [115] Srinivasa Rao G, Pethan Rajan N, Hari Sekhar M, Ammaji S, Chary KVR. Porous
of acrolein, 1995, US patent 5387720, (1995). zirconium phosphate supported tungsten oxide solid acid catalysts for the
[85] Neher A HTD, Process for the simultaneous production of 1,2- and 1,3- vapour phase dehydration of glycerol. J Mol Catal A: Chem 2014;395:486–93.
propanediol, 1995, US patent 5426249, (1995). [116] Gu Y, Liu S, Li C, Cui Q. Selective conversion of glycerol to acrolein over sup-
[86] Foo GS, Wei D, Sholl DS, Sievers C. Role of Lewis and Brønsted acid sites in ported nickel sulfate catalysts. J Catal 2013;301:93–102.
the dehydration of glycerol over niobia. ACS Catal 2014;4:3180–92.

Please cite this article as: A. Galadima, O. Muraza, A review on glycerol valorization to acrolein over solid acid catalysts, Journal of the
Taiwan Institute of Chemical Engineers (2016), http://dx.doi.org/10.1016/j.jtice.2016.07.019
JID: JTICE
ARTICLE IN PRESS [m5G;August 10, 2016;13:11]

A. Galadima, O. Muraza / Journal of the Taiwan Institute of Chemical Engineers 000 (2016) 1–16 15

[117] Znaiguia R, Brandhorst L, Christin N, Bellière Baca V, Rey P, Millet J-MM, [147] Ma T, Yun Z, Xu W, Chen L, Li L, Ding J, et al. Pd-H3PW12O40/Zr-MCM-41: an
et al. Toward longer life catalysts for dehydration of glycerol to acrolein. Mi- efficient catalyst for the sustainable dehydration of glycerol to acrolein. Chem
croporous Mesoporous Mater 2014;196:97–103. Eng J 2016;294:343–52.
[118] Cavani F, Guidetti S, Trevisanut C, Ghedini E, Signoretto M. Unexpected [148] Talebian-Kiakalaieh A, Amin NAS, Zakaria ZY. Gas phase selective conversion
events in sulfated zirconia catalyst during glycerol-to-acrolein conversion. of glycerol to acrolein over supported silicotungstic acid catalyst. J Ind Eng
Appl Catal, A 2011;409–410:267–78. Chem 2016;34:300–12.
[119] Ulgen A, Hoelderich WF. Conversion of glycerol to acrolein in the presence of [149] Talebian-Kiakalaieh A, Amin NAS. Supported silicotungstic acid on zirco-
WO3/TiO2 catalysts. Appl Catal, A 2011;400:34–8. nia catalyst for gas phase dehydration of glycerol to acrolein. Catal Today
[120] Lauriol-Garbey P, Postole G, Loridant S, Auroux A, Belliere-Baca V, Rey P, 2015;256:315–24 Part 2.
et al. Acid–base properties of niobium-zirconium mixed oxide catalysts for [150] Qiu F, Wang X, Zhang X, Liu H, Liu S, Yeung KL. Preparation and proper-
glycerol dehydration by calorimetric and catalytic investigation. Appl Catal, B ties of TS-1 zeolite and film using Sil-1 nanoparticles as seeds. Chem Eng
2011;106:94–102. J 2009;147:316–22.
[121] Cavani F, Guidetti S, Marinelli L, Piccinini M, Ghedini E, Signoretto M. The [151] Han W, Kwan SM, Yeung KL. Zeolite applications in fuel cells: water manage-
control of selectivity in gas-phase glycerol dehydration to acrolein catalysed ment and proton conductivity. Chem Eng J 2012;187:367–71.
by sulfated zirconia. Appl Catal, B 2010;100:197–204. [152] Chau JLH, Wan YSS, Gavriilidis A, Yeung KL. Incorporating zeolites in micro-
[122] Shen L, Yin H, Wang A, Lu X, Zhang C. Gas phase oxidehydration of chemical systems. Chem Eng J 2002;88:187–200.
glycerol to acrylic acid over Mo/V and W/V oxide catalysts. Chem Eng J [153] Chen X, Lam KF, Yeung KL. Selective removal of chromium from differ-
2014;244:168–77. ent aqueous systems using magnetic MCM-41 nanosorbents. Chem Eng J
[123] Lee YY, Lee KA, Park NC, Kim YC. The effect of PO4 to Nb2 O5 catalyst on the 2011;172:728–34.
dehydration of glycerol. Catal Today 2014;232:114–18. [154] Possato LG, Diniz RN, Garetto T, Pulcinelli SH, Santilli CV, Martins L. A com-
[124] García-Sancho C, Cecilia JA, Moreno-Ruiz A, Mérida-Robles JM, Santamaría– parative study of glycerol dehydration catalyzed by micro/mesoporous MFI
González J, Moreno-Tost R, et al. Influence of the niobium supported zeolites. J Catal 2013;300:102–12.
species on the catalytic dehydration of glycerol to acrolein. Appl Catal, B [155] Corma A. Inorganic solid acids and their use in acid-catalyzed hydrocarbon
2015;179:139–49. reactions. Chem Rev 1995;95:559–614.
[125] Dalil M, Carnevali D, Dubois J-L, Patience GS. Transient acrolein selectivity [156] Sheldon RA, Downing RS. Heterogeneous catalytic transformations for envi-
and carbon deposition study of glycerol dehydration over WO3 /TiO2 catalyst. ronmentally friendly production. Appl Catal, A 1999;189:163–83.
Chem Eng J 2015;270:557–63. [157] Egeblad K, Kustova M, Klitgaard SK, Zhu K, Christensen CH. Mesoporous ze-
[126] Cecilia J, García-Sancho C, Mérida-Robles J, González JS, Moreno-Tost R, olite and zeotype single crystals synthesized in fluoride media. Microporous
Maireles-Torres P. WO3 supported on Zr doped mesoporous SBA-15 silica for Mesoporous Mater 2007;101:214–23.
glycerol dehydration to acrolein. Appl Catal, A 2016;516:30–40. [158] Kim YT, Jung K-D, Park ED. A comparative study for gas-phase dehydration of
[127] Shen L, Yin H, Wang A, Feng Y, Shen Y, Wu Z, et al. Liquid phase dehydra- glycerol over H-zeolites. Appl Catal, A 2011;393:275–87.
tion of glycerol to acrolein catalyzed by silicotungstic, phosphotungstic, and [159] Rac V, Rakić V, Stošić D, Otman O, Auroux A. Hierarchical ZSM-5, Beta and
phosphomolybdic acids. Chem Eng J 2012;180:277–83. USY zeolites: acidity assessment by gas and aqueous phase calorimetry and
[128] Thomas A, Dablemont C, Basset JM, Lefebvre F. Comparison of H3PW12O40 catalytic activity in fructose dehydration reaction. Microporous Mesoporous
and H4SiW12O40 heteropolyacids supported on silica by 1H MAS NMR. C R Mater 2014;194:126–34.
Chim 2005;8:1969–74. [160] Boroń P, Chmielarz L, Gurgul J, Łatka ˛ K, Gil B, Marszałek B, et al. Influ-
[129] Bardin BB, Bordawekar SV, Neurock M, Davis RJ. Acidity of keggin-type ence of iron state and acidity of zeolites on the catalytic activity of FeHBEA,
heteropolycompounds evaluated by catalytic probe reactions, sorption mi- FeHZSM-5 and FeHMOR in SCR of NO with NH3 and N2 O decomposition. Mi-
crocalorimetry, and density functional quantum chemical calculations. J Phys croporous Mesoporous Mater 2015;203:73–85.
Chem B 1998;102:10817–25. [161] Almutairi SMT, Mezari B, Pidko EA, Magusin PCMM, Hensen EJM. Influence
[130] Shen L, Feng Y, Yin H, Wang A, Yu L, Jiang T, et al. Gas phase dehydration of of steaming on the acidity and the methanol conversion reaction of HZSM-5
glycerol catalyzed by rutile TiO2 -supported heteropolyacids. J Ind Eng Chem zeolite. J Catal 2013;307:194–203.
2011;17:484–92. [162] García-Trenco A, Martínez A. Direct synthesis of DME from syngas on hybrid
[131] Yan W, Suppes GJ. Low-pressure packed-bed gas-phase dehydration of glyc- CuZnAl/ZSM-5 catalysts: new insights into the role of zeolite acidity. Appl
erol to acrolein. Ind Eng Chem Res 2009;48:3279–83. Catal, A 2012;411–412:170–9.
[132] Sereshki BR, Balan SJ, Patience GS, Dubois JL. Reactive vaporization of crude [163] Yu S, Tian H. Acidity characterization of rare-earth-exchanged Y zeolite using
glycerol in a fluidized bed reactor. Ind Eng Chem Res 2010;49:1050–6. 31P MAS NMR. Chin J Catal 2014;35:1318–28.
[133] Suprun W, Lutecki M, Haber T, Papp H. Acidic catalysts for the dehydration [164] Milina M, Mitchell S, Michels N-L, Kenvin J, Pérez-Ramírez J. Interdependence
of glycerol: activity and deactivation. J Mol Catal A: Chem 2009;309:71–8. between porosity, acidity, and catalytic performance in hierarchical ZSM-5
[134] Liu L, Wang B, Du Y, Borgna A. Supported H4 SiW12O40/Al2 O3 solid acid zeolites prepared by post-synthetic modification. J Catal 2013;308:398–407.
catalysts for dehydration of glycerol to acrolein: evolution of catalyst [165] Groen JC, Moulijn JA, Pérez-Ramírez J. Decoupling mesoporosity formation
structure and performance with calcination temperature. Appl Catal, A and acidity modification in ZSM-5 zeolites by sequential desilication–dealu-
2015;489:32–41. mination. Microporous Mesoporous Mater 2005;87:153–61.
[135] Atia H, Armbruster U, Martin A. Influence of alkaline metal on performance [166] Mavrodinova V, Popova M, Mihályi MR, Pál-Borbély G, Minchev C. Influence
of supported silicotungstic acid catalysts in glycerol dehydration towards of the acidity modification on the reactions of toluene and ethyl benzene dis-
acrolein. Appl Catal, A 2011;393:331–9. proportionation. Effect of Lewis acidic InO+ species introduced into Y zeolite.
[136] Atia H, Armbruster U, Martin A. Dehydration of glycerol in gas phase using Appl Catal, A 2004;262:75–83.
heteropolyacid catalysts as active compounds. J Catal 2008;258:71–82. [167] Mihályi MR, Kollár M, Klébert S, Mavrodinova V. Transformation of ethylben-
[137] Kim YT, Jung KD, Park ED. Gas-phase Dehydration of glycerol over supported zene-m-xylene feed over MCM-22 zeolites with different acidities. Appl Catal,
silicotungstic acids catalysts. Bull Korean Chem Soc 2010;31:3283–90. A 2014;476:19–25.
[138] Gu Y, Liu S, Li C, Cui Q. Selective conversion of glycerol to acrolein over sup- [168] Epelde E, Aguayo AT, Olazar M, Bilbao J, Gayubo AG. Modifications in the
ported nickel sulfate catalysts. J Catal 2013;301:93–102. HZSM-5 zeolite for the selective transformation of ethylene into propylene.
[139] Possato LG, Diniz RN, Garetto T, Pulcinelli SH, Santilli CV, Martins L. A com- Appl Catal, A 2014;479:17–25.
parative study of glycerol dehydration catalyzed by micro/mesoporous MFI [169] Manjunathan P, Maradur SP, Halgeri AB, Shanbhag GV. Room temperature
zeolites. J Catal 2013;300:102–12. synthesis of solketal from acetalization of glycerol with acetone: effect of
[140] Lauriol-Garbey P, Millet JMM, Loridant S, Bellire-Baca V, Rey P. New efficient crystallite size and the role of acidity of beta zeolite. J Mol Catal A: Chem
and long-life catalyst for gas-phase glycerol dehydration to acrolein. J Catal 2015;396:47–54.
2011;281:362–70. [170] Epelde E, Gayubo AG, Olazar M, Bilbao J, Aguayo AT. Modified HZSM-5 zeo-
[141] Viswanadham B, Srikanth A, Kumar VP, Chary KVR. Vapor phase dehydration lites for intensifying propylene production in the transformation of 1-butene.
of glycerol to acrolein over SBA-15 supported vanadium substituted phospho- Chem Eng J 2014;251:80–91.
molybdic acid catalyst. J Nanosci Nanotechnol 2015;15:5391–402. [171] Wang Z, Wang L, Jiang Y, Hunger M, Huang J. Cooperativity of Brønsted and
[142] Kang TH, Choi JH, Bang Y, Yoo J, Song JH, Joe W, et al. Dehydration of glycerin Lewis acid sites on zeolite for glycerol dehydration. ACS Catal 2014;4:1144–7.
to acrolein over H3PW12O40 heteropolyacid catalyst supported on silica-alu- [172] Almutairi SMT, Mezari B, Filonenko GA, Magusin PCMM, Rigutto MS,
mina. J Mol Catal A: Chem 2015;396:282–9. Pidko EA, et al. Influence of extraframework aluminum on the Brønsted acid-
[143] Liu R, Wang T, Liu C, Jin Y. Highly selective and stable CsPW/Nb2O5 catalysts ity and catalytic reactivity of faujasite zeolite. ChemCatChem 2013;5:452–66.
for dehydration of glycerol to acrolein. Chin J Catal 2013;34:2174–82. [173] Li S, Zheng A, Su Y, Zhang H, Chen L, Yang J, et al. Brønsted/Lewis acid syn-
[144] Haider MH, Dummer NF, Zhang D, Miedziak P, Davies TE, Taylor SH, ergy in dealuminated HY zeolite: A Combined Solid-State NMR and Theoret-
et al. Rubidium- and caesium-doped silicotungstic acid catalysts supported ical Calculation Study. J Am Chem Soc 2007;129:11161–71.
on alumina for the catalytic dehydration of glycerol to acrolein. J Catal [174] Huang J, Jiang Y, Marthala VRR, Thomas B, Romanova E, Hunger M. Charac-
2012;286:206–13. terization and acidic properties of aluminum-exchanged zeolites X and Y. J
[145] Ning L, Ding Y, Chen W, Gong L, Lin R, Yuan L, et al. Glycerol dehydration Phys Chem C 2008;112:3811–18.
to acrolein over activated carbon-supported silicotungstic acids. Chin J Catal [175] Haw JF, Nicholas JB, Xu T, Beck LW, Ferguson DB. Physical organic chemistry
2008;29:212–14. of solid acids: lessons from in situ NMR and theoretical chemistry. Acc Chem
[146] Chai SH, Wang HP, Liang Y, Xu BQ. Sustainable production of acrolein: prepa- Res 1996;29:259–67.
ration and characterization of zirconia-supported 12-tungstophosphoric acid [176] Yoda E, Ootawa A. Dehydration of glycerol on H-MFI zeolite investigated by
catalyst for gas-phase dehydration of glycerol. Appl Catal, A 2009;353:213–22. FT-IR. Appl Catal, A 2009;360:66–70.

Please cite this article as: A. Galadima, O. Muraza, A review on glycerol valorization to acrolein over solid acid catalysts, Journal of the
Taiwan Institute of Chemical Engineers (2016), http://dx.doi.org/10.1016/j.jtice.2016.07.019
JID: JTICE
ARTICLE IN PRESS [m5G;August 10, 2016;13:11]

16 A. Galadima, O. Muraza / Journal of the Taiwan Institute of Chemical Engineers 000 (2016) 1–16

[177] Dalla Costa BO, Peralta MA, Querini CA. Gas phase dehydration of glycerol [200] Katryniok B, Paul S, Dumeignil F. Recent Developments in the field of cat-
over, lanthanum-modified beta-zeolite. Appl Catal, A 2014;472:53–63. alytic dehydration of glycerol to acrolein. ACS Catal 2013;3:1819–34.
[178] Chai SH, Wang HP, Liang Y, Xu BQ. Sustainable production of acrolein: in- [201] da Silva CXA, Mota CJA. The influence of impurities on the acid-catalyzed
vestigation of solid acid-base catalysts for gas-phase dehydration of glycerol. reaction of glycerol with acetone. Biomass Bioenergy 2011;35:3547–51.
Green Chem 2007;9:1130–6. [202] Hoang TQ, Zhu X, Danuthai T, Lobban LL, Resasco DE, Mallinson RG.
[179] Tao LZ, Yan B, Liang Y, Xu BQ. Sustainable production of acrolein: catalytic Conversion of glycerol to alkyl-aromatics over zeolites. Energy Fuels
performance of hydrated tantalum oxides for gas-phase dehydration of glyc- 2010;24:3804–9.
erol. Green Chem 2013;15:696–705. [203] Carriço CS, Cruz FT, dos Santos MB, Oliveira DS, Pastore HO, Andrade HMC,
[180] Nimlos MR, Blanksby SJ, Qian X, Himmel ME, Johnson DK. Mechanisms of et al. MWW-type catalysts for gas phase glycerol dehydration to acrolein. J
glycerol dehydration. J Phys Chem A 2006;110:6145–56. Catal 2016;334:34–41.
[181] Kim YT, Jung KD, Park ED. Gas-phase dehydration of glycerol over ZSM-5 cat- [204] Zhang H, Hu Z, Huang L, Zhang H, Song K, Wang L, et al. Dehydration of
alysts. Microporous Mesoporous Mater 2010;131:28–36. glycerol to acrolein over hierarchical ZSM-5 zeolites: effects of mesoporosity
[182] Decolatti HP, Dalla Costa BO, Querini CA. Dehydration of glycerol to acrolein and acidity. ACS Catal 2015;5:2548–58.
using H-ZSM5 zeolite modified by alkali treatment with NaOH. Microporous [205] Choi Y, Park H, Yun YS, Yi J. Effects of catalyst pore structure and acid prop-
Mesoporous Mater 2015;204:180–9. erties on the dehydration of glycerol. ChemSusChem 2015;8:974–9.
[183] Brand HV, Redondo A, Hay PJ. Theoretical studies of CO adsorption on [206] Lok BM, Messina CA, Patton RL, Gajek RT, Cannan TR, Flanigen EM. Silicoa-
H-ZSM-5 and hydrothermally treated H-ZSM-5. J Mol Catal A: Chem luminophosphate molecular sieves: another new class of microporous crys-
1997;121:45–62. talline inorganic solids. J Am Chem Soc 1984;106:6092–3.
[184] Fleischer U, Kutzelnigg W, Bleiber A, Sauer J. 1H NMR chemical shift and in- [207] Li Y, Huang Y, Guo J, Zhang M, Wang D, Wei F, et al. Hierarchical SAPO-34/18
trinsic acidity of hydroxyl groups. Ab initio calculations on catalytically active zeolite with low acid site density for converting methanol to olefins. Catal
sites and gas-phase molecules. J Am Chem Soc 1993;115:7833–8. Today 2014;233:2–7.
[185] Marques JP, Gener I, Ayrault P, Bordado JC, Lopes JM, Ribeiro FR, et al. In- [208] Yang G, Wei Y, Xu S, Chen J, Li J, Liu Z, et al. Nanosize-enhanced life-
frared spectroscopic study of the acid properties of dealuminated BEA zeo- time of SAPO-34 catalysts in methanol-to-olefin reactions. J Phys Chem C
lites. Microporous Mesoporous Mater 2003;60:251–62. 2013;117:8214–22.
[186] Guisnet M, Ayrault P, Coutanceau C, Alvarez MF, Datka J. Acid properties of [209] Aghaei E, Haghighi M. Effect of crystallization time on properties and cat-
dealuminated beta zeolites studied by IR spectroscopy. J Chem Soc - Faraday alytic performance of nanostructured SAPO-34 molecular sieve synthesized at
Trans 1997;93:1661–5. high temperatures for conversion of methanol to light olefins. Powder Tech-
[187] Armaroli T, Gutiérrez Alejandre A, Bevilacqua M, Trombetta M, Milella F, nol 2015;269:358–70.
Ramírez J, et al. 13-P-25-FTIR studies of the interaction of aromatic and [210] Aghamohammadi S, Haghighi M. Dual-template synthesis of nanostructured
branched aliphatic compounds with internal, external and extraframework CoAPSO-34 used in methanol to olefins: effect of template combinations on
sites of MFI-type zeolite materials. In: Galarneau FFFDRA, Vedrine J, editors. catalytic performance and coke formation. Chem Eng J 2015;264:359–75.
Studies in surface science and catalysis, 135. Elsevier; 2001. p. 346. [211] Aghaei E, Haghighi M. High temperature synthesis of nanostructured Ce-S-
[188] Jia C-J, Liu Y, Schmidt W, Lu A-H, Schüth F. Small-sized HZSM-5 zeolite as APO-34 catalyst used in conversion of methanol to light olefins: effect
highly active catalyst for gas phase dehydration of glycerol to acrolein. J Catal of temperature on physicochemical properties and catalytic performance. J
2010;269:71–9. Porous Mater 2015;22:187–200.
[189] Carriço CS, Cruz FT, Santos MB, Pastore HO, Andrade HMC, Mascaren- [212] Abdollahi S, Ghavipour M, Nazari M, Behbahani RM, Moradi GR. Effects of
has AJS. Efficiency of zeolite MCM-22 with different SiO2 /Al2 O3 molar ratios static and stirring aging on physiochemical properties of SAPO-18 and its per-
in gas phase glycerol dehydration to acrolein. Microporous Mesoporous Mater formance in MTO process. J Nat Gas Sci Eng 2015;22:245–51.
2013;181:74–82. [213] Li B, Tian P, Li J, Chen J, Yuan Y, Su X, et al. Synthesis of SAPO-35 molecular
[190] Santos Marques AL, Fontes Monteiro JL, Pastore HO. Static crystalliza- sieve and its catalytic properties in the methanol-to-olefins reaction. Chin J
tion of zeolites MCM-22 and MCM-49. Microporous Mesoporous Mater Catal 2013;34:798–807.
1999;32:131–45. [214] Aghaei E, Haghighi M. Enhancement of catalytic lifetime of nanostructured
[191] Derouane EG, Chang CD. Confinement effects in the adsorption of simple SAPO-34 in conversion of biomethanol to light olefins. Microporous Meso-
bases by zeolites. Microporous Mesoporous Mater 20 0 0;35-36:425–33. porous Mater 2014;196:179–90.
[192] Brändle M, Sauer J. Acidity differences between inorganic solids induced by [215] Qian Q, Ruiz-Martínez J, Mokhtar M, Asiri AM, Al-Thabaiti SA, Basahel SN,
their framework structure. a combined quantum mechanics/molecular me- et al. Single-catalyst particle spectroscopy of alcohol-to-olefins conversions:
chanics ab initio study on zeolites. J Am Chem Soc 1998;120:1556–70. Comparison between SAPO-34 and SSZ-13. Catal Today 2014;226:14–24.
[193] Biaglow AI, Gorte RJ, White D. 13C NMR studies of acetone in dealuminated [216] Álvaro-Muñoz T, Márquez-Álvarez C, Sastre E. Enhanced stability in the
faujasites: a probe for nonframework alumina. J Catal 1994;150:221–4. methanol-to-olefins process shown by SAPO-34 catalysts synthesized in
[194] Chai SH, Wang HP, Liang Y, Xu BQ. Sustainable production of biphasic medium. Catal Today 2013;215:208–15.
acrolein: gas-phase dehydration of glycerol over Nb2 O5 catalyst. J Catal [217] Sedighi M, Keyvanloo K. Kinetic study of the methanol to olefin process on a
2007;250:342–9. SAPO-34 catalyst. Front Chem Sci Eng 2014;8:306–11.
[195] Gu Y, Cui N, Yu Q, Li C, Cui Q. Study on the influence of channel structure [218] Chai S-H, Wang H-P, Liang Y, Xu B-Q. Sustainable production of acrolein: in-
properties in the dehydration of glycerol to acrolein over H-zeolite catalysts. vestigation of solid acid–base catalysts for gas-phase dehydration of glycerol.
Appl Catal, A 2012;429–430:9–16. Green Chem 2007;9:1130–6.
[196] Jia CJ, Liu Y, Schmidt W, Lu AH, Schüth F. Small-sized HZSM-5 zeolite as [219] Sarkar B, Pendem C, Konathala LS, Tiwari R, Sasaki T, Bal R. Cu nan-
highly active catalyst for gas phase dehydration of glycerol to acrolein. J Catal oclusters supported on nanocrystalline SiO2 –MnO2 : a bifunctional catalyst
2010;269:71–9. for the one-step conversion of glycerol to acrylic acid. Chem Commun
[197] Zhang L, Qiao S, Jin Y, Cheng L, Yan Z, Lu GQ. Hydrophobic functional group 2014;50:9707–10.
initiated helical mesostructured silica for controlled drug release. Adv Funct [220] Suprun W, Gläser R, Papp H. Gas-phase dehydration of glycerol over bifunc-
Mater 2008;18:3834–42. tional catalysts. Gas 2010;10:12.
[198] de Oliveira AS, Vasconcelos SJS, de Sousa JR, de Sousa FF, Filho JM, [221] Lourenco JP, Fernandes A, Bertolo RA, Ribeiro MF. Gas-phase dehydration of
Oliveira AC. Catalytic conversion of glycerol to acrolein over modified molec- glycerol over thermally-stable SAPO-40 catalyst. RSC Adv 2015;5:10667–74.
ular sieves: Activity and deactivation studies. Chem Eng J 2011;168:765–74. [222] Suprun W, Lutecki M, Haber T, Papp H. Acidic catalysts for the dehydration
[199] Kim YT, Jung K-D, Park ED. Gas-phase dehydration of glycerol over ZSM-5 of glycerol: activity and deactivation. J Mol Catal A: Chem 2009;309:71–8.
catalysts. Microporous Mesoporous Mater 2010;131:28–36.

Please cite this article as: A. Galadima, O. Muraza, A review on glycerol valorization to acrolein over solid acid catalysts, Journal of the
Taiwan Institute of Chemical Engineers (2016), http://dx.doi.org/10.1016/j.jtice.2016.07.019

You might also like