You are on page 1of 14

Journal of Sol-Gel Science and Technology

https://doi.org/10.1007/s10971-020-05344-6

ORIGINAL PAPER: SOL–GEL AND HYBRID MATERIALS FOR CATALYTIC,


PHOTOELECTROCHEMICAL AND SENSOR APPLICATIONS

Glycerol etherification towards selective diglycerol over mixed


oxides derived from hydrotalcites: effect of Ni loading
M. Aloui1 J. A. Cecilia

2 ●
R. Moreno-Tost2 S. B. Ghorbel1 M. Saïd Zina1 E. Rodríguez-Castellón2
● ● ●

Received: 22 January 2020 / Accepted: 12 June 2020


© Springer Science+Business Media, LLC, part of Springer Nature 2020

Abstract
A series of tetrametallic hydrotalcite with different Ni loading, were synthesized by coprecipitation method. Then, these
hydrotalcites were calcined in their respective metal oxides and were used as catalysts in the etherification reaction for the
synthesis of polyglycerols. The effect of Ni loading on the structural and textural features was investigated by various
techniques such as X-ray diffraction (XRD), thermogravimetric analysis TG–DSC–MS, temperature-programmed reduction
(H2-TPR), temperature-programmed desorption (NH3-TPD) and 1-butene isomerization as a model reaction to probe
1234567890();,:
1234567890();,:

acid–base character of catalysts. Mixed oxides derived from hydrotalcites are found to be active and suitable, via solvent
free, in the glycerol etherification reaction. A gradual enhance of glycerol conversion is revealed when increasing the Ni/Mg
molar ratio and the most active catalyst found is HTc-Ni75% with full selectivity to diglycerol.

Graphical Abstract

Keywords Hydrotalcite Etherification Glycerol Diglycerol Biodiesel


● ● ● ●

Supplementary information The online version of this article (https://


doi.org/10.1007/s10971-020-05344-6) contains supplementary
material, which is available to authorized users.

* J. A. Cecilia 2
Departamento de Química Inorgánica, Cristalografía y
jacecilia@uma.es
Mineralogía, Facultad de Ciencias, Universidad de Málaga,
1 Campus de Teatinos, 29071 Málaga, Spain
Laboratoire de Chimie des Matériaux et Catalyse, Faculté des
sciences de Tunis, Université de Tunis El Manar, Tunis 2092,
Tunisia
Journal of Sol-Gel Science and Technology

Highlights
● The presence of Ni2+ species modifies the amount of acid and basic sites.
● The partial incorporation of Ni2+ enhances the catalytic activity.
● The glycerol etherification leads to diglycerol and triglycerols as products.
● Temperatures higher than 220 °C lead to uncontrolled polymerizations.

1 Introduction etherification reaction. They found diglycerol selectivity


about 53% and glycerol conversion around 98%. From
The depletion of fossil fuels has caused the search and experimental and theoretical studies, it has been reported
development of alternative sources. Between them, the that basic sites are involved in the glycerol etherification
biomass is one of the most sustainable alternative since it reactions but the Lewis acid sites also have a clear influence
allows the obtaining of energy and valuable chemicals. One on the catalytic activity [11, 13], which has been confirmed
of the most mature strategies to obtain fuels is the synthesis by various authors [14–16]. In fact, solid acid catalysts such
of biodiesel where vegetable oils, like soybean oil, palm oil, as ionic exchange resins [17, 18], zeolites [17], clay
etc., and short-chain alcohols. The synthesis of biodiesel minerals [12, 19], or mesoporous silicas (SBA-15 and
also leads glycerol as by-product. The increasing biodiesel MCM-41) modified with aluminum species into the fra-
production has led to a surplus of glycerol, causing a mework [16] have shown to be active in glycerol ether-
decrease in the price since in many cases is considered as ification reaction.
invaluable waste. Layered double hydroxides (LDHs), also named hydro-
Glycerol originated from biodiesel industry can be con- talcites (HTs) and represented by the general formulae
xþ n x
verted into polyglycerols by an etherification reaction in the ½M2þ
1x Mx ðOHÞ2  ½Ax=n 

 mH2 O, are very promising
presence of a basic catalyst and without solvent. Poly- materials in several fields such as adsorbent material or in
glycerols have an enormous potential in wide variety of catalysis. The calcination of LDHs between 450 and 600 °C
fields such as polymers, cosmetics, food, dispersants, leads to formation of mixed oxides with Lewis basic sites
pharmaceutical industries, lubricants, biomedical, or drug [20, 21]. The use of LDHs as catalytic precursor has a high
delivery systems [1–4]. potential to replace the homogenous-based catalysts since
The conversion of glycerol to reach desired short-chain they are easy to prepare, relatively inexpensive catalysts,
polyglycerol remains a scientific challenge, in this sense the environmentally benign, and recyclable [22].
design of suitable catalysts to control the selectivity to Recently, the mixed oxides obtained from thermal
specific products is highly desired [5]. First studies were treatment for the LDHs have been used as promising cata-
focused on the use of basic homogeneous catalysts like lysts in the etherification reaction. Thus, Garcia-Sancho
hydroxides, carbonates, and oxides [6, 7] as well as acidic et al., have evaluated the influence of the type of precipitant
homogeneous catalysts [8, 9] (such as sulfuric, benzene- agent, the calcination temperature in the specific surface
sulfonic, and dodecyl benzene-sulfonic acids). In fact, these area and the dispersion of MgO–Al2O3 species for the
catalysts recognize a high solubility in glycerol and/or a glycerol etherification to obtain polyglycerols [23]. On
nonselective oligomerization because of the generation of the other hand, Pérez-Barrado et al. [14] have analyzed the
high molecular weight glycerol oligomers coupled to others influence of the acid–base properties of calcined MgAl and
by-products through dehydration reaction [10]. Thus, the CaAl layered double hydroxide to obtain short-chain
attention has been shifted to the use of heterogeneous cat- polyglycerols. Lately, Guerrero-Urbaneja et al. [24] have
alysts, which are easily recovered, more suitable, and optimized the Mg/Fe molar ratio of the LDHs to obtain
selective for short-chain polyglycerol. short-chain polyglycerols, reaching a conversion around
Ruppert et al. [11] investigated the use of alkaline-earth 25% with diglycerol selectivity higher than 90% in all
metal oxides, like MgO, CaO, SrO, and BaO. These authors cases. In addition, LDHs have been synthesized with a wide
pointed out that glycerol conversion is directly related to the variety of mono- (Cu+) [25] and tetravalent metal cations
amount and strength of the basic sites, reaching a high (Zr4+) [26, 27] to be used as catalytic precursor since these
glycerol conversion around 60%, with selectivity to di- and modifications improve the physicochemical properties of
triglycerol higher than 90%. However, alkaline-earth oxides the final catalysts.
suffer from partial hydroxylation by water formed in the The present research modifies the acid–base properties of
etherification reaction, making them instable at high tem- the metal mixed oxide coming from LDHs by a substitution
perature reaction (220–260 °C). Ayoub et al. [12] studied of Al species by Zr species. Previous research has pointed
montmorillonite K-10 with intercalated LiOH for selective out that metal mixed oxide obtained by calcining MgAl-
production of diglycerol via solvent free glycerol LDH display high proportion of Lewis acid sites on its
Journal of Sol-Gel Science and Technology

surface, which can be attributed to the presence of Al3+ 2.3 Samples characterization
species. The modification of Al3+ by other species with
Lewis sites, such as Zr4+ can affect to the nature, amount, Powder diffraction patterns were collected on an PANalytical
and strength of these acid sites. On the other hand, the automated diffractometer, EMPYREAN model, using Cu-
Lewis acidity of the Mg2+ species can be modified by the Kα_1,2 (1.5406 Å) radiation and a last generation PIXel
substitution of these Mg2+ by Ni2+ species. Thus, the aim of detector. The database employed were ICSD, PDF 2-2004,
this research is the synthesis of Mg/Ni/Al/Zr HTs with and COD of High ScorePlus programme of PANalytical. The
different Ni loading by coprecipitation method in order to crystallite size (D) was calculated using the Williamson–Hall
improve their catalytic properties for the target reaction. equation [28], B cosθ = (K λ/D) + (2 ε sinθ), where θ is the
Structural and textural features of mixed oxide derived from Bragg angle, B is the full width at half maximum (FWHM)
hydrotalcite with different Ni loadings were studied and of the XRD peaks, K is the Scherrer constant, λ is the
catalytic performances of all mixed oxides were evaluated wavelength of the X ray, and ε the lattice strain [28].
in the glycerol etherification reaction. Thermogravimetric analysis (TGA) and differential
scanning calorimetry (DSC) were performed using a ther-
mal Analyzer from METTLER TOLEDO model TGA/DSC
2 Experimental 1. Analysis were performed in Al2O3 crucible 70 µl in the
temperature range 30–900 °C, with nitrogen flow (50 ml/
2.1 Chemicals min) and a heating rate of 10 °C min−1. The equipment was
coupled to a PFEIFFER VACUUM ThermoStar TM GSD
Precursor salt used in this work to synthesize the LDHs 320 model mass spectrometer. The mass ratio measurement
were magnesium nitrate hexahydrate, Mg(NO3)2 6H2O range was varied from 1 to 200. It consists of a Faraday
(>98%, Sigma-Aldrich), aluminum nitrate nonahydrate, Al detector and a SEM detector. The capillary was made in
(NO3)2 9H2O (>98%, Sigma-Aldrich), zirconium oxyni- quartz and was operated at 200 °C. The scanning speed was
trate hydrate, ZrO(NO3)2 H2O (99%, Sigma-Aldrich), 2–60 uma/s, depending on the type of scan performed. The
and nickel nitrate hexahydrate Ni(NO3)2 6H2O (>98%, vacuum with a working pressure in the order of 10−6 mbar
Sigma-Aldrich). Sodium hydroxide, NaOH (99%, Sigma- is achieved by a membrane pump and a turbo.
Aldrich), and sodium carbonate Na2CO3 (99.5%, Norma- The textural parameters were determined from the N2
pur) were used as precipitating agents. Finally, glycerol adsorption–desorption isotherms at −196 °C, obtained by
(98%, Sigma-Aldrich) was used as fed for the etherifica- using an automatic ASAP 2020 model of gas adsorption
tion reaction. analyzer supplied for Micromeritics. Prior to N2 adsorption,
the sample were evacuated at 450 °C (heating rate 10 °C
2.2 Catalysts preparation min−1) for 18 h. Surface areas were determined by using the
Brunauer–Emmet–Teller (BET) equation, considered a N2
Hydrotalcite-like compounds, were synthesized at room molecule cross section of 16.2 Å2 [29]. The total pore
temperature by the coprecipitation method at low super- volume was calculated from the adsorption isotherm at P/
saturation, where M2+/(Al3++Zr4+) global ratio was main- P0 = 0.996. The density functional theory method was
tained at 2, Al3++Zr4+ was at 0.33 in all cases, while the employed to determine the pore size distribution [30].
Ni2+/M2+ molar ratio was varied from 25 until 100%. The The acidity of catalyst was determined by temperature-
total solution concentration of M2+ + Al3+ + Zr4+ was 1 M. programmed desorption of ammonia (NH3-TPD). In each
To synthesize LDHs, a basic solution of NaOH/Na2CO3 experiment, 80 mg of the calcined LDH were activated from
with a [OH−] concentration of 2 M and [CO32−] con- room temperature to 550 °C in He flow (35 ml min−1),
centration of 0.5 M in an aqueous solution of metal salts maintaining the temperature 2 h. After that, the sample was
was added to achieve a pH of 10. The resulting gel was cooled to 100 °C and exposed to pure NH3 flow for 5 min.
aged at 60 °C for 18 h, then the gel was filtered, washed Later, the physisorbed-NH3 was removed using He flow
with deionized water to neutral pH, and dried in the oven at (35 ml min−1) and then, the physisorbed ammonia was
80 °C. All catalysts were activated in a tubular furnace at removed using a He flow (35 ml min−1). Finally, the NH3-
450 °C for 15 h, under a flow of helium to obtain the TPD was carried out from 100 °C to 550 °C, maintaining
derived mixed oxide. this temperature for 15 min using similar He flow as carrier
Mixed oxides with different nickel loads were designated gas. The NH3 desorbed was registered using a thermal
as follow: HTc-Ni25%, HTc-Ni50%, HTc-Ni75%, and HTc- conductivity reactor (TCD).
Ni100%, where 25, 50, 75, and 100% indicates the Ni2+ The catalytic isomerization of 1-butene was performed in
content in wt.% of the M2+ species. a tubular glass flow microreactor. Overall, 80 mg of
Journal of Sol-Gel Science and Technology

calcined LHDs were pretreated for 2 h in a He flow at microfiltered to analyze the amount of Mg2+, Ni2+, Al3+,
450 °C (30 ml min−1), and experiments were carried out at and Zr4+ in the reaction medium by IPC-AES.
400 °C temperature. Experiments were performed in a
tubular glass flow microreactor at τ = 0.6 gcat g−11-but h, and 2.4 Reaction procedure
the time on stream was 120 min. The reactant 1-butene and
the reaction products were analyzed on-line in a gas chro- The catalytic activity was evaluated in the etherification of
matograph (Shimadzu GC-14B) equipped with a wide bore glycerol between 210 and 240 °C in a three-necked glass
KCl/AlCl3 column. For this reaction, the distribution of batch reactor without solvent, equipped with a water-
n-butene was close to the equilibrium point. cooled condenser coupled with a Dean–Stark system to
The basicity was studied by temperature-programmed remove the formed water, thermometer, and vigorous
desorption of CO2. Overall, 100 mg of calcined LDH was stirring. The atmosphere of the reactor was kept inert by
pretreated under a He stream at 450 °C for 2 h (2 °C min−1, means of a N2 flow of 15 mL min−1. Before the reaction,
100 mL min−1). Then, the sample was cooled until 100 °C, 300 mg of sample was activated at different temperatures
and a flow of pure CO2 (50 mL min−1) was subsequently for 15 h (heating rate, 2 °C min−1), under a helium flow.
introduced into the reactor during 1 h. The CO2-TPD was The reactions were carried out with a loading charge of
carried out between 100 and 800 °C under a helium flow 2–6 wt.%, a temperature of 210–240 °C and a reaction time
(10 °C min−1, 30 mL min−1), and evolved CO2 was ana- of 1–24 h. After the reaction, the reaction medium was
lyzed by an on-line with a TCD, after passing by an ice- cooled and separated by filtration on a porous plate. The
NaCl trap to eliminate any trace of water. reaction products were analyzed by means of gas chro-
The reducibility of Ni species was determined by H2- matography. An aliquot (ca 80 mg) of the reaction mixture
thermoprogrammed reduction (H2-TPR). Overall, 80 mg of was dissolved in dried pyridine (Aldrich) and then N,O-bis
calcined LDH were previously treated with a He flow (trimethylsilyl)trifluoroacetamide was added. This solution
(35 mL min−1) at 450 °C for 2 h. After cooling to room was aged in a stove at 60 °C during 1 h and analyzed in a
temperature, the H2 consumption was studied between this gas chromatograph (Shimadzu GC model 14A) equipped
temperature and 800 °C, by using an Ar/H2 flow (48 mL with FID and a capillary silica fused TBR-14 column
min−1, 10 vol.% of H2) and a heating rate of 10 °C min−1. (Teknokroma). The only detected products were unreacted
Water formed in the reduction process was removed with an glycerol, di-, and triglycerol. The selectivities to the dif-
isopropanol-liquid nitrogen trap and a cold finger (−80 °C). ferent products were calculated as the weight ratio of the
The H2 consumption was measured with an on-line TCD, respective product to the sum of products formed. The
and quantified by calibration with pure CuO as reference glycerol conversion was calculated as the ratio of the
standard (Aldrich), assuming a total reduction of CuO to detected glycerol to the sum of the glycerol and products
Cu0. formed.
X-ray photoelectron spectroscopy (XPS) studies were The detected products were unreacted glycerol, di-, and
performed with a Physical Electronics PHI 5700 spectro- triglycerols. The glycerol conversion was determined from
meter equipped with a hemispherical electron analyzer (Eq. 1):
(model 80-365B) and a Mg Kα (1253.6 eV) X-ray source. n0  nf
High-resolution spectra were recorded at 45° take-off-angle xGly ¼  100; ð1Þ
n0
by a concentric hemispherical analyzer operating in the
constant pass energy mode at 29.35 eV, using a 720-μm where, n0 is the initial amount of glycerol mol and nf is the
diameter analysis area. Charge referencing was done against final amount of glycerol mol.
adventitious carbon (C 1s at 284.8 eV). The pressure in the The Selectivity (Solig) was calculated using (Eq. 2).
analysis chamber was kept lower than 5 × 10−6 Pa. PHI nolig
ACCESS ESCA-V6.0 F software package was used for data Solig ¼  100; ð2Þ
n0  nf
acquisition and analysis. A Shirley-type background was
subtracted from the signals. Recorded spectra were always where nolig is the amount of each oligomer in mol.
fitted using Gauss–Lorentz curves in order to determine Yields were calculated using (Eq. 3).
more accurately the binding energy of the different element
XGly Σ Solig
core levels. Yolig ¼ : ð3Þ
100
The leaching of the mixed oxides was evaluated by ICP-
AES by using a Perkin Elmer (model ELAN DRC-e) Oligoglycerols higher than triglycerol were not detected by
spectrometer. To analyze the leaching of the catalysts, the this analytical procedure, so that other products not detected
reaction liquid was dissolved with H2O (liquid/H2O ratio were labeled as others, which could include acrolein or
1:9 (v/v)) to diminish its viscosity. This liquid was then glycidol among others.
Journal of Sol-Gel Science and Technology

Between each run, the liquid was mixed with H2O, (110) reflection (a = 2d110) (Table 1). The c parameter is
establishing a ratio reaction liquid/H2O of 1:9 (v/v) to three times the distance between adjacent hydroxide and it
decrease its viscosity. Then, the solid was filtered and dried has been also calculated from the (003) reflection (c =
to be used in the next run. 3d003). The obtained values are very close to those reported
by other authors [31, 32] for HTs with carbonates in the
interlayer space, which confirms that the incorporation of
3 Results and discussion Ni2+ and Zr4+ species does not cause clear modifications in
the lamellar structure of the HTs.
3.1 Catalyst characterization The increase of Ni content has led to the formation of
wider diffractions peaks, which suggests a progressive loss of
3.1.1 Powder X-ray diffraction the crystallinity. Considering that both cations present similar
ionic radii, 0.52 Å Mg2+ cation and 0.72 Å for Ni2+ cation,
The XRD patterns of the as-prepared HTs containing Ni the isomorphic substitution can cause changes in the 2θ
with different Ni2+/M2+ molar ratio are shown in Fig. 1. values of the diffractions peaks, being only noticeable a
XRD data of the synthesized HTs displayed the typical slight displacement at lower 2θ values only for the sample
reflections peaks of a well-crystallized HT (JCPDS file: 70- HT-Ni100%. The gradual modification in the width and
2151) belonging to the space group R 3 m and related to the intensity of the diffraction peaks when Mg2+ is replaced by
presence of the (003), (006), and (009) crystal planes in the Ni2+ species could be attributed to both cations precipitate at
layered structure. The diffractograms discard the segrega- different pH values, which can affect the self-assembly of the
tion of phases of Zr4+ species, indicating that Zr4+ cations lamellar structure. The crystal size was estimated by the
are incorporated into the structure or at least are well dis- Williamson–Hall method [28], obtaining crystals with a size
persed. In addition, the absence of diffraction peaks attrib- lower than 5 nm. In fact, the progressive substitution of Mg2+
uted to nickel species also indicates a good incorporation of by Ni2+ in the synthetic step causes a slight decrease in the
Ni2+ cations into the brucite layers. The unit cell parameter, crystal size.
a, is the average distance between two metal ions within the
layers, and it has been calculated from the d-spacing of the 3.1.2 Thermogravimetric (TG)/DSC analysis and evolved CO2

In order to determinate the optimum calcination temperature


of the LDHs TG and DSC curves as well as the evolved
CO2 followed by mass spectroscopy (m/z = 44) were
investigated. From TG curves (Supplementary Information,
Fig. 1A), two main typical regions are observed. The first
region occurred below 250 °C, corresponding to the loss of
physisorbed and interlayer water with no significant influ-
ence on the HT-lamellar structure. On the other hand, the
second region arise between 250 and 800 °C, related to the
dehydroxylation of the brucite sheets by water removal and
decarboxylation by decomposition of carbonate ions,
resulting in the collapse of the layer structure and the for-
mation of carbon dioxide. In addition, it is known that the
thermal stability of HTs depends on the nature of the layer
cations and interlayer anions, comparing the TG/DSC
curves of all samples (Supplementary Information, Fig. 1B),
Fig. 1 XRD patterns of hydrotalcites: a HT-Ni25%, b HT-Ni50%, c HT- it is found that the variation of Ni content led to some
Ni75%, and d HT-Ni100% changes in the thermal stability of these compounds. Both

Table 1 XRD and TG results of Catalyst M2+/Al3++Zr4+ Parameter a (Å ) Parameter c (Å ) Cristallites size (nm) Total weight losses (%)
fresh hydrotalcite-like
compounds HT-Ni25% 2.00 3.05 23.43 4.1 43.7
HT-Ni50% 2.00 3.04 23.41 3.9 40.0
HT-Ni75% 2.00 3.03 23.38 3.5 40.8
HT-Ni100% 2.00 3.04 23.47 3.2 37.7
Journal of Sol-Gel Science and Technology

Fig. 2 XRD patterns of hydrotalcite calcined at 450 °C

TG and DSC analysis reveals that LDHs with lower nickel


content lose higher proportion of physisorbed and interlayer Fig. 3 TEM micrographs of mixed oxides: a HTc-Ni25%, b HTc-Ni50%,
water (Table 1). c HTc-Ni75%, and d HTc-Ni100%
With regard to the second step, DSC analysis reveals as
Ni loading increased the temperature of this second peak is
decreased. This fact was confirmed following the CO2 attributed to Al2O3 or ZrO2 in any case, confirming both the
emission signal (m/z = 44) as illustrated in Supplementary low crystallinity of these oxides and the incorporation of
Information, Fig. 2, where the maximum value is shifted Al3+ and Zr4+ in the structure of MgO.
from 375 °C for HT-Ni25% to 320 °C for HT-Ni100%, which
could be related to a slight decrease of the crystallinity for 3.1.3 TEM micrograph
higher Ni content and with a lower interaction between the
brucite layer with the CO32− species used to counterbalance The morphology of the LDHs thermally treated was eval-
the charge deficiency. Considering the DSC analysis, each uated by TEM (Fig. 3). It is noteworthy that all micrographs
LDH was calcined at 450 °C for 2 h to obtain the corre- display homogenous structures, discarding the formation of
sponding mixed oxide. segregated species with higher particle size, as was sug-
The thermal treatment under helium flow leads to the gested by XRD (Fig. 2). TEM micrographs show that the
destruction of the hydrotalcite structure with remarkable mixed oxide particles are formed by aggregates of small
structural changes and the formation of mixed oxide phase. structures with laminar or fibrous forms. In addition, it is
The diffraction patterns (Fig. 2) of all mixed oxides are noteworthy that an increase of the nickel content led to the
similar and present low crystallinity with broad peaks. The formation of smaller aggregates.
main peaks characteristic of hydrotalcite structure dis-
appeared, and new diffraction peaks appeared about 36°, 3.1.4 N2 adsorption–desorption at −196 °C
42°, and 63°. Taking into account that these diffractions
peaks could be ascribed to NiO or MgO species, it is The textural properties of mixed oxides obtained by thermal
impossible to discern between both phases since both treatment of LDHs were determined from the N2
phases display the same packaging and similar ionic radius adsorption–desorption at −196 °C (Supplementary Infor-
(Mg2+ = 0.52 Å and Ni2+ = 0.72 Å) so the diffraction peaks mation, Fig. 3). According to the IUPAC classification, all
of each phase are very close so it is not possible to deter- isotherms can be considered as type-II, which are typical of
mine the size of each oxide specie. However, the diffraction macroporous materials [33], as indicates the increase of the
peak at 2θ = 36° increases the intensity as the Ni loading N2 adsorbed at higher relative pressures. Taking into
does. This issue could indicate about some segregation of account that the mixed oxides itself lack porosity, the spe-
NiO. In any case, the diffraction peaks of the mixed oxides cific surface area SBET is attributed to the N2-filled in the
are very broad, suggesting the presence of a small particle space between the small layers. In any case, all oxides
size and high dispersion. The XRD profiles of the calcined showed similar SBET values from 194 to 227 m2/g, which
samples do not reflect the existence of diffraction peaks means a high dispersion of these mixed oxides. The
Journal of Sol-Gel Science and Technology

hysteresis loop can be adjusted to type H3 hysteresis for all 3.1.6 XPS results
calcined HTs, which is usually found in materials consisting
of aggregates of particles forming slit-shaped pores with In order to determinate the chemical composition of the
nonuniform size or shape [33]. mixed oxides, XPS measurements were carried out. The C
1s core level spectra display two contributions located at
3.1.5 Reducibility and reduction temperature of catalysts 284.8 eV and 289.4–289.7 eV, which are ascribed to the
presence of adventitious carbon and carbonate species,
The study of the reducibility of nickel species was investi- respectively. In spite the thermal treatment at 450 °C, it is
gated to evaluate how the dispersion or the interaction with noteworthy the existence of carbonate species. In this sense,
other oxide species changes when the nickel loading is previously other authors have also detected carbonate spe-
modified. The data is shown in Supplementary Information, cies after the thermal treatment of the LDHs [24], being
Fig. 4, reveal that the reduction of all mixed oxides takes attributed to the remaining strong bonded carbonates and/or
place in a step, and could be assigned to the reduction of Ni2+ the recarbonation of the mixed oxides during the sample
species in mixed oxide to Ni0, in the range 300–900 °C. handling Table 2.
However, the width of this peak reveals the existence of The O 1s core level spectra also show the coexistence of
different Ni2+ species interacting with different metals. It two bands located at 529.8–530.1 eV and 531.5–531.9 eV,
seems that as Mg2+ loading is decreased (i.e., the Ni2+ is which are assigned to the presence of O2− and species like
increased) the H2-consumption takes place at lower tem- –OH or carbonates groups [24]. From the data obtained in
perature, so the Ni–Mg interaction should be stronger, which Table 2, it can be observed that the concentrations of both
is in agreement with the literature [34]. contributions are similar in all cases, although the propor-
tion of oxides species slightly increases in the case of HTc-
Ni100%.
In the case of the Mg 2p core level spectra, it is only
noteworthy the existence of a single contribution located
between 49.5 and 49.9 eV, which is assigned to Mg2+ in an
oxidic environment. It is also noticeable that the Mg con-
tent decreases according as Mg2+ species are replaced by
Ni2+ ones, as was expected. The analysis of the theoretical
Ni2+/M2+ reveals that the obtained values are higher than
the theoretical ones, which indicates an enrichment of the
surface with Ni2+ for the mixed oxides. As Mg2+ and Ni2+
displays different ionic radii, the phase segregation cannot
be discarded since the diffraction peaks of both phases
appear at similar 2θ values. Thus, it is possible that a
portion of Ni2+ can be segregated during the thermal
treatment, located mainly on the surface, as the atomic
concentrations points out (Table 2). Both Al 2p and
Zr 3d core level spectra showed unique contribution loca-
ted at 73.5–74.0 eV in the case of Al 2p region and
Fig. 4 Profiles of temperature-programmed desorption of ammonia for 182.2–182.5 eV for the Zr 3d region, assigned in all cases
hydrotalcites calcined at 450 °C to oxides species. With regard to the Zr4+/(Al3+ + Zr4+)

Table 2 XPS results of the mixed oxides coming from hydrotalcites


M 2þ Zr 4þ Ni2þ Ni2þ
Sample Binding energy, eV (atomic concentrations,%) M 3þ þM 4þ M 3þ þM 4þ M 2þ Theoretical M 2þ

C 1s O 1s Mg 2p Al 2p Ni 2p Zr 3d

HTc-Ni25% 284.8 (3.7) 530.1 (39.8) 49.9 (15.7) 74.0 (9.0) 855.9 (7.3) 183.5 (3.2) 1.89 0.26 0.32 0.25
289.5 (2.9) 531.9 (18.5)
HTc-Ni50% 284.8 (2.1) 529.9 (38.0) 49.5 (10.0) 73.9 (9.4) 855.8 (14.8) 183.4 (2.8) 2.02 0.22 0.59 0.50
289.7 (1.8) 531.7 (21.1)
HTc-Ni75% 284.8 (1.7) 529.9 (37.9) 49.4 (5.7) 73.5 (9.8) 855.5 (22.0) 183.5 (2.9) 2.18 0.22 0.97 0.75
289.4 (1.2) 531.5 (18.8)
HTc-Ni100% 284.8 (4.7) 529.8 (43.6) – 73.5 (6.6) 855.4 (26.3) 183.2 (2.2) 2.98 0.25 1.00 1.00
289.4 (1.9) 531.7 (14.7)
Journal of Sol-Gel Science and Technology

ratio, the data shown in Table 2 report that those ratios are Table 3 Amount of acid sites, basic sites, and H2 adsorption over
mixed oxides catalysts
slightly lower than the theoretical value (0.33), indicating
that Zr4+ are less exposed on the surface of the mixed Catalyst NH3 CO2 Basic/ Hydrogen
oxides. Finally, the (Ni2+ + Mg2+)/(Al3+ + Zr4+) molar adsorbed adsorbed acid ratio adsorption
(µmol/g) (µmol/g) (mmol H2/g)
ratio is close to 2.00 in all samples, which is in agreement
with the theoretical data, except in HTc-Ni100% where the HTc-Ni25% 358 112 0.31 0.6
proportion of M2+ species is higher on the surface because HTc-Ni50% 310 118 0.38 11.6
of the higher segregation of oxide phases. HTc-Ni75% 396 117 0.29 19.5
HTc-Ni100% 136 67 0.49 23.0
3.1.7 Acid–base properties
W weak acid site strength, M moderate strength, S strong strength

As was indicated previously, both acid and basic sites are


involved in the etherification reactions so the quantification enrichment leads to a decrease of the acid sites at higher
of these active sites may become a key parameter in the temperatures which could be ascribed to the Al3+ or Zr4+
evaluation of the catalytic behavior. cations, more electronegative than Ni2+. Also, the intensity
The amount and strength of acid sites in calcined HTs of the peak at low temperature is lowered as Ni2+ replaces
were investigated by temperature-programmed desorption Mg2+, therefore we can tentatively ascribe those acid sites
of ammonia (NH3-TPD). The NH3-TPD profiles of all cal- to Mg2+ species.
cined HTs (Fig. 4) show a broad asymmetrical shape indi- Shen et al. [36] admitted the presence of both Brønsted and
cating the presence of acid sites showing different strengths. Lewis acid site in HTs calcined in similar temperature range.
Previously, some authors have considered the deconvolu- Brønsted sites are related to surface protons and are weaker
tion of the NH3-TPD profile in three desorption sites [35]. than Lewis sites, which are associated with Al3+ cations
Weak acid sites occurred at temperature below 200 °C, located in octahedral sites [37]. The incorporation of Zr4+
moderate acid sites located in the temperature range of cations in the synthetic step also provides Lewis sites [38].
200–350 °C and strong acid sites strength associated to The total numbers of acid sites of catalysts, illustrated in
temperatures above 350 °C. Table 3, reveals that HTc-Ni75% is the sample with the highest
As the Ni2+ loading some effects are observed from the amount of total acid sites in comparison with others catalysts.
TPD-NH3 plots. First, the HTc-Ni100% only shows a broad As was indicated previously, the NH3-TPD is an
band of desorption of ammonia, leading to only the pre- appropriate technique to evaluate the strength and amount
sence of weak–moderate of acid strength. When Mg and Ni of acid sites; however, it is impossible to discern between
are both present in the hydrotalcite, the mixed oxides dis- acid and basic sites. 1-Butene isomerization reaction is
play two visible bands. The band at low temperature considered a model reaction to the acidity of these calcined
decreases the intensity as the Ni loading increases and the HTs [39, 40]. This reaction proceeds by acid centers and
band at higher temperature is shifted to lower temperatures takes place through a first step where the C = C bond is
as the Ni/Mg ratio increases. Taking into account that all the moved, obtaining cis- and trans-2-butene, as products. In a
solids present the same amounts of Al and Zr, these features second step, the isomerization to obtain isobutene takes
could be ascribed to the relationship between Ni and Mg. place, although this step requires the presence of stronger
The cationic radii of Ni is of 0.72 Å and the corresponding acid sites. The use of catalysts with amount and strong acid
to Mg is of 0.52 Å, this could imply that when the Ni sites can cause the cracking reactions with the formation of
loading is low it could be incorporated to the periclase C2, C3, C5, and C6 olefins, the dimerization to octanes or
structure and therefore it is well dispersed and it generates even the formation of carbonaceous deposits. 1-butene
Lewis acid sites with moderate–strong strength of acid sites isomerization is an interesting reaction since the type of
and as Ni/Mg ratio is increased this peak decreases the acidity and the strength of the active sites affect the selec-
intensity and is shifted to lower temperatures. So, this peak tivity strongly. There are two types of isomerization. The
could be ascribed to acidic Ni2+ species. Also, the peak at double bond isomerization does not require very strong
low temperature is lowered its intensity as Ni/Mg increases, acidity. In this case, the cis/trans-2-butenes ratio provides
therefore those acid sites could be ascribed to Mg2+ species. insights into the acid (cis/trans < 1) or basic (cis/trans > 1)
As the Ni2+ loading is increased, the broad band at high nature of the catalyst. The other isomerization reaction is
temperature is shifted to lower temperatures and finally attributed to methyl shift, which only proceeds in strong
when all Mg2+ cations are substituted by Ni2+, the curve is sites. The dehydrogenation of 1-butene is directly related
a broad band ranging from 200 to 350 °C. As it was men- with the Ni content. In this sense, previous research has
tioned above, the XPS analysis showed an enrichment of reported that NiO-based catalysts favor the dehydrogenation
the catalyst surface by nickel (probably as NiO) so this reactions [41].
Journal of Sol-Gel Science and Technology

Table 4 Activity and selectivity


Catalyst Conversion of Cis-2- Trans-2- Isobutane 1,3-butadiene Cis/trans ratio
in 1-butene isomerization for all
1-butene butene butene
mixed oxides
HTc-Ni25% 34.8 58.3 36.7 2.9 2.1 1.59
HTc-Ni50% 31.6 57.3 37.1 3.1 2.5 1.54
HTc-Ni75% 46.3 54.0 40.0 2.6 3.3 1.35
HTc-Ni100% 26.1 42.8 34.4 11.9 10.8 1.24

Fig. 5 Profiles of temperature-programmed desorption of CO2 for


hydrotalcites calcined at 450 °C Fig. 6 Diglycerols yield for different catalysts

The data shown in Table 4 reveal that the conversion in [42, 43]. CO2 desorption data, shown in Fig. 6, reveal that
the 1-butene isomerization reaction is relatively low the desorption located at higher temperature decreases
(between 31.6 and 46.3%) for the mixed oxides Mg–Ni. according as Mg2+ are substituted by Ni2+, and also is
This value decreases in the case of calcined hydrotalcite shifted to lower desorption temperatures while the deso-
without Mg, which only reaches a conversion value of rption located at lower temperature is maintained. These
6.1%. With regard to the selectivity, both cis- and trans-2- data suggest that oxide ions of the Mg species (Mg2+–O2−)
butene are the main products in all cases. The analysis of the provide stronger basic sites, while the basicity attributed to
cis-/trans-2-butene ratio reveals values above than one in all NiO must be weaker.
cases, which confirms the basic nature of the catalyst. The
selectivity towards isobutene and 1,3-butadiene defines the 3.2 Catalytic activity
strength of the acid sites [39, 40]. Data shown in Table 3
indicate that HTc-Ni25%, HTc-Ni50%, and HTc-Ni75% exhibit 3.2.1 Performance of Ni/Mg molar ratio on the catalytic
a low proportion of acid sites with high strength. In the case activity
of HTc-Ni100%, the activity in the 1-butene isomerization is
lower, although the selectivity to isobutene and 1,3-buta- The HTc-Ni mixed oxides with different Ni loading were
diene is slightly higher, which supposes low proportion of investigated in glycerol etherification using a batch reactor
acid sites than those calcined HTs containing magnesium. between 210 and 240 °C, at atmospheric pressure under N2
The amount and strength of the basic sites were deter- atmosphere.
mined by CO2-TPD (Fig. 5). In all cases, it can be observed In a preliminary test, the thermal contribution was eval-
the presence of two well defined bands, which indicates the uated in the glycerol etherification reaction in the absence of
existence basic sites with different strength. Previous catalyst (Fig. 6). The catalytic data after 24 h at 220 °C
research has reported that the first CO2 desorption located at show a glycerol conversion of only 10% with a diglycerol
lower temperature, in the range of 100–300 °C, is attributed selectivity of 93%. With regard to the mixed oxides, all
to the desorption of bicarbonate formed on weakly OH samples are active for the etherification reaction of glycerol
groups, while the second broad CO2 desorption band, which with full selectivity towards diglycerol. The catalytic data
takes place between 450 and 700 °C, is related with the reveals that a gradual enhance of the diglycerol yield from
unidentated carbonated basic formed with surface O2− ions 29% for HTc-Ni25% to 55% for HTc-Ni75%. A commercial
Journal of Sol-Gel Science and Technology

hydrotalcite, supplied by Aldrich, was calcined and used as the basic sites and to increase the amount of available acid
reference in the synthesis of polyglycerols using similar sites, MgAl HTs were synthesized and then calcined to
catalytic conditions, reaching a glycerol conversion of 21%, obtain the mixed oxides [23]. These mixed oxides reached
being only detected diglycerols. The data also were com- a highest conversion value of 50.7% with selectivity
pared with the MgO prepared by activation of the corre- towards diglycerols of 84.8% under similar reaction
sponding nitrate at 700 °C, obtaining a glycerol conversion conditions, were used in this study [23]. In the same way,
of 31% with a diglycerol selectivity of 90% [13], while the these authors also synthesized MgFeOx mixed oxides
Al2O3 synthesized by similar synthetic procedure only from HTs as active phase in the glycerol etherification
reaches a diglycerol yield close to 10%. obtaining a glycerol conversion of 41% with selectivity
The catalytic data, shown in Fig. 6, indicates that the close to 90% [24]. Considering the catalytic data reported
presence of Ni2+ species has a positive effect in the catalytic in the literature using similar reaction conditions, it seems
behavior of the glycerol etherification. In this sense, pre- clear that both the incorporation of Ni2+ species into the
vious research pointed out that the incorporation of Ni2+ hydrotalcite improve the glycerol conversion, obtaining
also caused an improvement of the catalytic behavior in diglycerols as unique product.
basic reactions such as the glycerol carbonate synthesis by The mixed oxide that exhibited the highest conversion
the presence of basic sites with variable strength [44]. It is values and diglycerol yield, i.e., HTc-Ni75%, was chosen to
well reported in the literature that the glycerol etherification optimize the reaction parameters in the glycerol etherification.
is a reaction that implies both basic and Lewis acid sites
[11, 13]. Data reported in Tables 3, 4, and Fig. 4 show how 3.2.2 Influence of the catalyst loading
the catalytic behavior follows similar trend to NH3-TPD and
1-butene isomerization from the HTc-Ni25% to HTc-Ni75%. First, the effect of the catalyst loading was evaluated on
In the case of the HTc-Ni100% the amount of the acid sites the glycerol etherification reaction at 220 °C and 24 h
diminishes along with the conversion of 1-butene. With (Fig. 7). Catalytic data reveals that the maximum con-
regard to the basic sites determined by CO2-TPD, it can be version value is reached using only 2 wt.% of HTc-Ni75%
observed that the amount of basic sites decreased for the catalyst, achieving a glycerol conversion of 55% with a
HTc-Ni100%, mainly by the absence of stronger basic sites. diglycerol selectivity of 100%. The increase of the cata-
The decrease of the amount both acid and basic sites can be lyst content does not seem to be beneficial neither for
related with the lower glycerol conversion in the HTc- glycerol conversion nor for diglycerol selectivity. Indeed,
Ni100% catalyst. the use of a loading charge of 6 wt.% causes a significant
In all cases, the conversion values are lower than those decrease of the diglycerol selectivity to reach 54% along
obtained using Na2CO3 as catalyst, which reaches a total with the formation of triglycerol product (STriglycerol =
conversion of 100% after 24 h of reaction [45]. In this 46%). The use of larger amount of catalyst can cause
sense, some authors have previously pointed out that the diffusional problems limiting the glycerol conversion. On
carbonate of alkaline metals show the highest glycerol the other hand, other authors have established that the
conversion due to the solubility of such carbonates in gly- decrease of glycerol conversion may occur because of the
cerol. However, these homogenous catalysts are not very back-scission of diglycerol to glycerol [46]. In addition,
selective for diglycerol [23, 45]. Nonetheless, the use of
heterogeneous catalysts also displays some advantages
related to the easier separation of the catalyst and the reu-
sability of the catalyst.
It is difficult to carry out a comparison between the
catalytic data previously reported in the literature with
those shown in the present research. Ruppert et al. eval-
uated the catalytic behavior of different alkaline-earth
metal oxides (MgO, CaO, SrO, and BaO) establishing that
the catalytic activity is directly related with the strength of
basic sites [11]. The use of catalysts with strong basic
sites causes the uncontrolled polymerization of the gly-
cerol. Barros et al., evaluated the catalytic activity of the
calcined eggshell (CaO) under similar catalytic condi-
tions, reaching higher conversion values [42], although
the selectivity towards diglycerols is lower due to the Fig. 7 Effect of catalyst loading on glycerol conversion, diglycerol,
stronger basic sites. In order to diminish the strength of and triglycerol selectivity for HTc-Ni75% catalyst
Journal of Sol-Gel Science and Technology

Fig. 8 Effect of reaction temperature on glycerol conversion, digly- Fig. 9 Diglycerols yield for HTc-Ni75% catalyst
cerols, triglycerols, and tetraglycerols selectivities for HTc-Ni75%
catalyst
3.2.4 Evaluation of time on stream
the use of larger amount of catalyst also causes an
increment of the amount of acid and basic sites and The influence of the reaction time in the glycerol ether-
therefore an increase of its reactivity, resulting in the reac- ification for the mixed oxides with highest conversion
tion evolves beyond diglycerol molecule due to uncon- values HTc-Ni75% was evaluated along 24 h at 210 °C using
trolled etherification. 2 wt.% of catalyst (Fig. 9). The catalytic data show how
both glycerol conversion and diglycerols selectivity
3.2.3 Influence of the reaction temperature increase as the reaction time does. It is noteworthy that the
conversion value is very low in the early hours, reaching a
The influence of the reaction temperature on the catalytic glycerol conversion of only 10% after 6 h at 210 °C. This
behavior of HTc-Ni75% was another parameter evaluated in fact suggests that it is necessary an activation step to favor
the glycerol etherification. The catalytic data reported in the etherification since a diglycerols yield of 55% is reached
Fig. 8 show how the glycerol conversion rises as the reac- after 24 h of reaction at 210 °C. In this sense, several
tion temperature increases from 50% after 24 h at 210 °C to authors have proposed a mechanism where the glycerol
61% after 24 h at 240 °C. Despite the higher conversion etherification takes place in two steps. First, the basic cen-
values with the temperature, the use of more severe conditions ters extract a hydrogen of the glycerol molecule and then an
also causes uncontrolled etherification, as indicates the unsaturated site attacks the deprotonated glycerol to another
decrease of the diglycerol selectivity according as the tem- glycerol molecule forming the diglycerol species [11, 13].
perature increases from 100 to 63% at 240 °C due to an Considering these data and those reported in the literature, it
uncontrolled etherification forming undesired products, which is expected that one of these steps of the reaction must be
are difficult to separate. Previous researches have pointed out limited by the temperature.
that the products with low polymerization-degree, i.e., di- and Another key point in the use of heterogeneous catalysts
triglycerols, have high industrial interest due to their wide is related to the leaching of the active phase. Previous
range of uses and applications [15, 47] while the formation of investigations have pointed out that La, Cs, or Mg-based
oligoglycerols due to larger polymerization are less interesting catalysts supported on mesoporous silica suffer from the
due to their physicochemical properties and the difficult to loss of active phase by leaching [48, 50]. However, the
separate the mixture of compounds [46]. In addition, the use concentration of Mg, Ni, Al, and Zr in the reaction medium
of higher temperature can favor the existence of undesired site can be considered negligible [51], confirming the high
reactions, which can generate toxic product such as acrolein stability of the mixed oxides in the reaction medium.
or glycidol. Under the reaction conditions reported in the
present study, these products have not been detected even 3.2.5 Reuse of the catalysts
[48, 49]. In this sense, several authors [23, 24, 42, 48] have
established that the optimum reaction temperature to obtain The reuse of the catalysts is a key parameter to attain sus-
diglycerols using solid basic catalyst between 200 and tainable and competitive catalysts (Fig. 10). This study was
220 °C. This temperature does not provide the highest gly- carried out using 2 wt.% of HTc-Ni75% catalyst at 220 °C for
cerol conversion; however, the side reactions are avoided, 24 h. Between each run, the catalyst was separated by
obtaining diglycerols as single product. decantation and washed with H2O. The catalytic data
Journal of Sol-Gel Science and Technology

with low crystal size, being evaluated in the selective


etherification of glycerol to yield mainly diglycerols. The
role of the Ni species has not been reported in the literature
for this reaction yet. The catalytic data reveals that the
incorporation of Ni species causes an increase of the acid
and basic sites, maintaining a SBET relatively high. This
fact is directly related with the catalytic activity in the
glycerol etherification.
The catalytic data show a highest glycerol conversion
of 55%, with full selectivity towards diglycerols after 24 h
of reaction at 210 °C with only 2 wt.% of HTc-Ni75%
catalyst. The increase of the catalyst loading does not
improve the catalytic behavior probably due to diffusional
limitations. An increase of the reaction temperature favors
Fig. 10 Diglycerols yield for HTc-Ni75% catalyst after several cycles the formation of triglycerols even tetraglycerols due to
higher temperature leads to uncontrolled polymerization,
obtaining products with a lower interest industry. Finally,
reveals that the glycerol conversion decrease from 55 to the mixed oxides have shown high stability since the
35% after 4 runs. In all cases, the single product obtained in leaching is practically negligible.
these reaction conditions was diglycerol.
The decrease of glycerol conversion could be ascribed to Acknowledgements This work was supported by the projects RTI2018-
several parameters. As was indicated previously, the 099668-B-C22 (Spanish Ministry of Science, Innovation and Uni-
versities, Spain and FEDER Funds) and UMA18-FEDERJA, P12-
leaching of the Mg, Ni, Al, or Zr can be considered as RNM-1565 projects (Junta de Andalucía, Spain). A special acknowledge
negligible so the homogeneous contribution in the glycerol to the Tunisian Ministry of Higher Education, Scientific Research,
etherification should be discarded. The washing of the used Information and Communication Technologies for the financial support.
catalyst with H2O facilities the removal of the reagent and
products strongly adsorbed. However, this washing can Compliance with ethical standards
hydroxylate the mixed oxides in such a way the acid-based
properties of the catalyst can be modified. The hydroxyla- Conflict of interest The authors declare that they have no conflict of
tion must cause a decrease of the acid and basic sites, which interest.
are involved in the glycerol etherification. Despite the
Publisher’s note Springer Nature remains neutral with regard to
decrease in catalytic activity after each run, washing with jurisdictional claims in published maps and institutional affiliations.
H2O seems to be the most appropriate technique to reuse the
catalyst since the separation by decantation results in a References
glycerol conversion of 21% while calcination gives rise to a
conversion of a 37% in the 2nd run. This more pronounced 1. Wilms D, Stiriba SE, Frey H (2010) Hyperbranched polyglycerols:
decrease can be ascribed to the decantation is not enough to from the controlled synthesis of biocompatible polyether polyols to
separate both reagents and products, which interacts multipurpose applications. Acc Chem Res 43:129–141
2. Calderón M, Quadir MA, Sharma SK, Haag R (2010) Dendritic
strongly with the mixed oxides species in such a way these polyglycerols for biomedical applications. Adv Mater 22:190–218
active sites are partially blocked by subsequent runs. In the 3. Salehpour S, Dube MA (2011) Towards the sustainable produc-
case the calcination method, the organic species adsorbed tion of higher-molecular-weight polyglycerol. Macromol Chem
on the surface of the catalyst are removed by its decom- Phys 212:1284–1293
4. Yadav JD, Kulkarni PR, Vaidya KA, Shelke GT (2011) Nio-
position at high temperature. This combustion process is an somes: a review. J Pharm Res 4:632–636
exothermic reaction so this treatment at high temperature 5. Zhou CH, Beltramini JN, Fan YX, Lu GQ (2008) Chemoselective
can cause sintering of the catalyst, which is directly related catalytic conversion of glycerol as a biorenewable source to
with a decrease of the amount of available active sites. valuable commodity chemicals. Chem Soc Rev 37:527–549
6. Richter M, Krisnandi YK, Eckelt R, Martin A (2008) Homo-
geneously catalyzed batch reactor glycerol etherification by
CsHCO3. Catal Commun 9:2112–2116
4 Conclusions 7. Krisnandi YK, Eckelt R, Schneider M, Martin A, Richter M
(2008) Glycerol upgrading over zeolites by batch-reactor liquid-
phase oligomerization: heterogeneous versus homogeneous reac-
A series of tetrametallic hydrotalcite with different Ni tion. ChemSusChem 1:835–844
loading, were synthesized by coprecipitation method. 8. Medeiros MA, Araujo MH, Augusti R, de Oliveira LCA, Lago RMJ
Then, these HTs were calcined obtaining mixed oxides (2009) Acid-catalyzed oligomerization of glycerol investigated by
Journal of Sol-Gel Science and Technology

electrospray ionization mass spectrometry. Braz Chem Soc 27. Velu S, Suzuki K, Okazaki M, Kapoor MP, Osaki T, Ohashi F
20:1667–1673 (2000) Oxidative steam reforming of methanol over CuZnAl(Zr)-
9. Harvey SB, Shen S (1996) One phase production of polyglycerol Oxide catalysts for the selective production of hydrogen for fuel
esters. Patent US 5585506A. cells: catalyst characterization and performance evaluation. J Catal
10. Lemke D, Nivens S (2007) Process for preparing polyglycerol and 194:373–384
mixed ethers. Patent WO 2007/092407. 28. Williamson GK, Hall WH (1953) X-ray line broadening from filed
11. Ruppert AM, Meeldijk JD, Kuipers BWM, Erné BH, Weckhuysen aluminium and wolfram. Acta Metall 1:22–31
BM (2008) Glycerol etherification over highly active CaO-based 29. Brunauer S, Emmett PH, Teller E (1938) Adsorption of gases in
materials: new mechanistic aspects and related colloidal particle multimolecular layers. J Am Chem Soc 60:309–319
formation. Chem Eur J 14:2016–2024 30. Landers J, Gor GY, Neimark AV (2013) Density functional theory
12. Ayoub M, Abdullah AZ (2013) LiOH-modified montmorillonite methods for characterization of porous materials. Colloid Surf. A
K-10 as catalyst for selective glycerol etherification to diglycerol. 437:3–32
Catal Commun 34:22–25 31. Aramendia MA, Avilés Y, Benitez JA, Borau V, Jiménez C,
13. Calatayud M, Ruppert AM, Weckhuysen BM (2009) Theoretical Marinas JM, Ruiz JR, Urbano FJ (1999) Comparative study of
study on the role of surface basicity and Lewis acidity on the Mg/Al and Mg/Ga layered double hydroxides. Micropor Mesopor
etherification of glycerol over alkaline earth metal oxides. Chem Mater 29:319–328
Eur J 15:10864–10870 32. Noda Pérez C, Pérez CA, Henriques CA, Monteiro JLF (2004)
14. Pérez-Barrado E, Pujol MC, Aguiló M, Llorca J, Cesteros Y, Díaz Hydrotalcites as precursors for Mg,Al-mixed oxides used as cat-
F, Pallarès J, Marsal LF, Salagre P (2015) Influence of acid–base alysts on the aldol condensation of citral with acetone. Appl Catal
properties of calcined MgAl and CaAl layered double hydroxides A 272:229–240
on the catalytic glycerol etherification to short-chain poly- 33. Thommes M, Kaneko K, Neimark AV, Olivier JP, Rodriguez-
glycerols. Chem Eng J 264:547–556 Reinoso F, Rouquerol J, Sing KSW (2015) Physisorption of gases,
15. Sivaiah MV, Robles-Manuel S, Valange S, Barrault J (2012) with special reference to the evaluation of surface area and pore
Recent developments in acid and base-catalyzed etherification of size distribution (IUPAC Technical Report). Pure Appl Chem
glycerol to polyglycerols. Catal Today 198:305–313 87:1051–1069
16. González MD, Salagre P, Mokaya R, Cesteros Y (2014) Tuning 34. Dębek R, Motak M, Duraczyska D, Launay F, Gálvez ME,
the acidic and textural properties of ordered mesoporous silicas for Grzybek T, da Costa P (2016) Methane dry reforming over
their application as catalysts in the etherification of glycerol with hydrotalcite-derived Ni–Mg–Al mixed oxides: the influence of Ni
isobutene. Catal Today 227:171–178 content on catalytic activity, selectivity and stability. Catal Sci
17. Cottin K, Clacens JM, Pouilloux Y, Barrault J (1998) Preparation Technol 6:6705–6715
of diglycerol and triglycerol by the direct polymerization of gly- 35. Sundaramurthy V, Dalai AK, Djaye JA (2008) The effect of
cerol in the presence of the new solid catalysts. Ol Corps Gras phosphorus on hydrotreating property of NiMo/γ-Al2O3 nitride
Lipide 5:407–412 catalyst. Appl Catal A 335:204–210
18. Richter M, Eckett R, Krisnandi YK, Martin A (2008) Verfahren 36. Shen J, Tu M, Hu C (1998) Structural and surface acid/base
zur selektiven herstellung von linearem diglycerin. Chem Ing properties of hydrotalcite-derived MgAlO oxides calcined at
Tech 80:1573–1577 varying temperatures. J Solid State Chem 137:295–301
19. Ayoub M, Abdullah AZ (2015) Diglycerol synthesis via solvent- 37. Lino AVP, Assaf EM, Assaf JM (2016) Hydrotalcites derived
free selective glycerol etherification process over lithium-modified catalysts for syngas production from biogas reforming: effect of
clay catalyst. Chem Eng J 225:784–789 nickel and cerium load. Catal Today 289:78–88
20. Tichit D, Gérardin C, Durand R, Coq B (2006) Layered double 38. García-Sancho C, Moreno-Tost R, Mérida-Robles J, Santamaría-
hydroxides: precursors for multifunctional catalysts. Top Catal González J, Jiménez-Lopez A, Maireles-Torres P (2012) Zirco-
39:89–96 nium doped mesoporous silica catalysts for dehydration of gly-
21. Xu ZP, Zhang J, Adebajo MO, Zhang H, Zhou CH (2011) Cat- cerol to high added-value products. Appl Catal A 433:179–187
alytic applications of layered double hydroxides and derivatives. 39. Patrono P, La Ginestra A, Ramis G, Busca G (1994) Conversion
Appl Clay Sci 53:139–150 of 1-butene over WO3-TiO2 catalysts. Appl Catal A 107:249–266
22. Fan G, Li F, Evans DG, Duan X (2014) Catalytic applications of 40. Cecilia JA, García-Sancho C, Franco F (2013) Montmorillonite
layered double hydroxides: recent advances and perspectives. based porous clay heterostructures: influence of Zr in the structure
Chem Soc Rev 43:7040–7066 and acidic properties. Micropor Mesopor Mater 176:95–102
23. García-Sancho C, Moreno-Tost R, Mérida-Robles JM, Santa- 41. Solsona B, Concepción P, López Nieto JM, Dejoz A, Cecilia JA,
maría-González J, Jiménez-López A, Maireles Torres P (2011) Agouram S, Soriano MD, Torres V, Jiménez-Jiménez J, Rodrí-
Etherification of glycerol to polyglycerols over MgAl mixed guez Castellón E (2016) Nickel oxide supported on porous clay
oxides. Catal Today 167:84–90 heterostructures as selective catalysts for the oxidative dehy-
24. Guerrero-Urbaneja P, García-Sancho C, Moreno-Tost R, Merida- drogenation of ethane. Catal Sci Technol 6:3419–3429
Robles J, Santamaria-Gonzalez J, Jimenez-Lopez A, Maireles- 42. Barros FJS, Moreno-Tost R, Cecilia JA, Ledesma-Muñoz AL, de
Torres P (2014) Glycerol valorization by etherification to poly- Oliveira LCC, Luna FMT, Vieira RS (2017) Glycerol oligomers
glycerols by using metal oxides derived from MgFe hydrotalcites. production by etherification using calcined eggshell as catalyst.
Appl Catal A 470:199–207 Mol Catal 433:282–290
25. Kuhl S, Tarasov A, Zander S, Kasatkin I, Behrens M (2014) Cu- 43. Correia LM, Campelo NS, Novaes DS, Cavalcante CL, Cecilia
based catalyst resulting from a Cu, Zn, Al hydrotalcite-like JA, Rodríguez-Castellón E, Vieira RS (2015) Characterization and
compound: a microstructural, thermoanalytical, and in-situ XAS application of dolomite as catalytic precursor for canola and
study. Chem Eur J 20:3782–3792 sunflower oils for biodiesel production. Chem Eng J 269:35–43
26. Gao P, Xie R, Wang H, Zhong L, Xia L, Zhang Z, Wei W, Sun Y 44. Liu P, Derchi M, Hensen EJM (2014) Promotional effect of
(2015) Cu/Zn/Al/Zr catalysts via phase-pure hydrotalcite-like transition metal doping on the basicity and activity of calcined
compounds for methanol synthesis from carbon dioxide. J CO2 hydrotalcite catalysts for glycerol carbonate synthesis. Appl Catal
Util 11:41–48 B144:135–143
Journal of Sol-Gel Science and Technology

45. Clacens JM, Pouilloux Y, Barrault J, Linares C, Goldwasser M (1998) lanthanum oxide supported on MCM-41: a heterogeneous basic
Mesoporous basic catalysts: comparison with alkaline exchange zeo- catalyst. Appl Catal A 479:76–86
lites (basicity and porosity). Application to the selective etherification 49. Ayoub M, Khayoon MS, Abdullah AZ (2012) Synthesis of oxy-
of glycerol to polyglycerols. Stud Surf Sci Catal 118:895–902 genated fuel additives via the solventless etherification of glycerol.
46. Martin A, Richter M (2011) Oligomerization of glycerol a critical Bioresource Technol 112:308–312
review. Eur J Lipid Sci Technol 113:100–17 50. Clacens JM, Pouilloux Y, Barrault J (2002) Selective etherifica-
47. Sutter M, da Silva E, Duguet N, Raoul Y, Métay E, Lemaire M tion of glycerol to polyglycerols over impregnated basic MCM-41
(2015) Glycerol ether synthesis: a bench test for green chemistry type mesoporous catalysts. Appl Catal A 227:181–190
concepts and technologies. Chem Rev 115:8609–8651 51. Barros FJS, Cecilia JA, Moreno-Tost R, de Oliveira MF, Rodríguez-
48. Gholami Z, Abdullah AZ, Lee KT (2014) Heterogeneously cata- Castellón E, Luna FMT, Vieira RS (2018) Glycerol oligomerization
lyzed etherification of glycerol to diglycerol over calcium- using low cost dolomite catalyst. Waste Biomass Valor 11:1499–1512

You might also like