You are on page 1of 28

Basic catalysis on MgO: generation,

characterization and catalytic properties


of active sites
J. I. Di Cosimo,* V. K. Dı́ez, C. Ferretti and
C. R. Apesteguı́a*
Published on 24 February 2014 on http://pubs.rsc.org | doi:10.1039/9781782620037-00001

DOI: 10.1039/9781782620037-00001

The generation, characterization and catalytic properties of MgO active sites were studied.
MgO samples stabilized at different temperatures were used to control the distribution of
surface base sites; specifically, MgO was calcined at 673 K, 773 K and 873 K (samples
MgO-673, MgO-773 and MgO-873). The nature, density and strength of MgO base sites
were characterized by temperature-programmed desorption of CO2 and infrared spec-
troscopy after CO2 adsorption at 298 K and sequential evacuation at increasing tem-
peratures. MgO samples contained surface sites of strong (low coordination O2 anions),
medium (oxygen in Mg2þ-O2 pairs) and weak (OH groups) basicity. The density of
strong basic sites was predominant on MgO-673. The increase of the calcination tem-
Downloaded on 16/02/2015 11:50:03.

perature drastically decreased the density of strong base sites and to a lesser extent that of
weak OH groups, while slightly increased that of medium-strength base sites. The
catalytic properties of MgO samples were proved for the aldol condensation of citral with
acetone to yield pseudoionone, the hydrogen transfer reaction of mesityl oxide with
2-propanol to obtain the unsaturated alcohol 4-methyl-3-penten-2ol, and the synthesis
of monoglycerides via the transesterification of methyl oleate with glycerol. The effect of
calcination temperature on the MgO catalytic properties depended on the basicity re-
quirements for the rate-limiting step of the base-catalyzed reaction. The activity for both
the aldol condensation of citral with acetone and the glycerolysis of methyl
oleate diminished with the MgO calcination temperature because these reactions
were essentially promoted on strongly basic O2 sites. In contrast, the synthesis of
4-methyl-3-penten-2ol by the hydrogen transfer reduction of mesityl oxide with
2-propanol increased with calcination temperature because the reaction intermediate
was formed on medium-strength Mg2þ-O2 pair basic sites. Additional information on the
role played by the MgO active sites on the kinetics of base-catalyzed reactions was
obtained by performing molecular modeling studies on our MgO catalysts using Density
Functional Theory (DFT) for the glycerolysis of methyl oleate, an unsaturated fatty acid
methyl ester (FAME). The molecular modeling of glycerol and FAME adsorptions was
carried out using terrace, edge and corner sites for representing the MgO (100) surface. In
agreement with catalytic results, calculations predicted that dissociative chemisorption of
glycerol with O–H bond breaking occurs only on strong base sites (edge sites) whereas
nondissociative adsorption takes place on medium-strength base sites such as those of
terrace sites. Results also indicated that glycerol was more strongly adsorbed than FAME.
The glycerol/FAME reaction would proceed then through a mechanism in which the most
relevant adsorption step is that of glycerol.

1 Introduction
Alkaline earth metal oxides catalyze a variety of organic reactions
requiring the cleavage of a C–H bond step and the formation of carba-
nion intermediates. In particular, pure and alkali-promoted MgO has

Catalysis Science and Engineering Research Group (GICIC), INCAPE,


UNL-CONICET. Santiago del Estero 2654. (3000) Santa Fe, Argentina.
E-mail: capesteg@fiq.unl.edu.ar; dicosimo@fiq.unl.edu.ar

Catalysis, 2014, 26, 1–28 | 1



c The Royal Society of Chemistry 2014
View Online

been shown to promote Cannizzaro and Tischenko reactions [1, 2],


Michael, Wittig and Knoevenagel condensations [3, 4], transesterifica-
tion reactions [5–8], double-bond isomerizations [9], self- and cross-
condensation reactions [10–13], Henry reaction [14], alcohol coupling
[15–17], and H2 transfer reactions [18]. However, the MgO basicity needed
for efficiently promoting these reactions depend on the rate-limiting step
requirements. MgO can be synthesized in a variety of presentation
Published on 24 February 2014 on http://pubs.rsc.org | doi:10.1039/9781782620037-00001

formats, including nanosheets [19], nanowires [20] and nanoparticles


[21], but its catalytic properties depend greatly on the preparation
method. Nevertheless, most of reports on the preparation of magnesia
deal with the effect of the synthesis method and conditions on the
MgO structural and physical properties [22–24]. Very few papers have
attempted to tailor the distribution, density, and strength of surface base
sites of MgO upon synthesis in order to design the catalyst surface to
reaction requirements [25–27]. More insight on the relationship between
the synthesis procedure with the generation and control of MgO surface
base sites is then required to improve the efficient use of this oxide in
catalysis applications.
Downloaded on 16/02/2015 11:50:03.

Detailed characterization of MgO base sites is crucial to establish


correlations between the surface basic properties and the catalyst activity
and selectivity for a given reaction. The most common methods for
characterization of solid basicity are thermal programmed desorption
(TPD) and infrared spectroscopy (IR) of preadsorbed probe molecules,
and the use of test reactions. TPD studies provide information on the
density and strength of base sites while additional insight on the base
site nature is often obtained by IR characterization. Carbon dioxide has
been largely employed as a probe molecule for evaluating the solid
basicity by TPD and IR techniques [28–31] although other acid molecules
such as acetic acid have been also used [32]. On the other hand, the test
reactions most frequently used for characterizing the catalyst acid-base
properties are the decomposition of alcohols, in particular 2-propanol
[33–35], 2-butanol [36, 37] and 2-methyl-3-butyn-2-ol [38–40]. In the case
of 2-propanol, it is generally accepted that 2-propanol dehydration
to propylene occurs on solid acids containing Brønsted acid sites via an
E1 mechanism while on amphoteric oxides with acid-base pair sites
propylene is obtained through a concerted E2 mechanism [41]. On strong
basic catalysts, 2-propanol is dehydrogenated to acetone via an E1cB
anionic mechanism [42]. Thus, the catalyst acid-base properties may be
related to the propylene/acetone selectivity ratio. In contrast, test
reactions have been used only in few cases for characterizing base site
strength distributions on solid bases. For example, in a previous work
[43], we proposed that on alkali-modified MgO catalysts 2-propanol
decomposition to acetone and propylene takes place via an E1cB
mechanism in two parallel pathways sharing a common 2-propoxy
intermediate; in this mechanism, the intermediate-strength base sites
promote acetone formation, whereas high-strength base sites selectively
yield propylene. Nevertheless, several studies have shown that the use of
test reactions is not sensitive enough to establish a basicity scale of the
catalysts [44].

2 | Catalysis, 2014, 26, 1–28


View Online

Theoretical calculations of surface sites have been performed for


exploration of MgO catalysis. In general, Density Functional Theory (DFT)
calculations have shown to be a powerful tool to characterize the thermal
stability of hydrated oxide surfaces [45]. Regarding MgO catalysts, DFT
studies on the structure of MgO surface defects have been carried out to
establish the stability of surface OH groups for water and methanol
adsorptions [46, 47]. Recently, combined IR and DFT studies have been
Published on 24 February 2014 on http://pubs.rsc.org | doi:10.1039/9781782620037-00001

performed in an attempt to specify the actual structure of the CO2 species


adsorbed on magnesium oxide surface [48]. Unfortunately, theoretical
calculations to predict the relationship between the basic site nature and
strength and the reaction mechanism have been done only for limited
cases.
In this work we study the generation, characterization and catalytic
properties of active sites on MgO catalysts. The base properties of MgO
samples obtained from Mg(OH)2 decomposition were tuned by modify-
ing the solid calcination temperature. The density and strength of MgO
surface base sites were determined by TPD and IR spectroscopy of CO2
adsorbed at 298 K. The activity and selectivity of MgO samples were probed
Downloaded on 16/02/2015 11:50:03.

for the liquid-phase cross-aldol condensation of citral with acetone to


obtain pseudoionones, the liquid-phase transesterification of methyl ole-
ate with glycerol to yield monoglycerides, and the gas-phase hydrogen
transfer reduction of mesityl oxide with 2-propanol toward 4-methyl-3-
penten-2ol. Besides, we performed DFT calculations to obtain additional
information on the role played by the MgO active sites on the kinetics of
base-catalyzed reactions. Specifically, we present molecular modeling
studies on our MgO catalysts for the glycerolysis of methyl oleate.

2 Experimental
2.1 Catalyst preparation
Magnesium oxide samples were prepared by hydration with distilled
water of low-surface area commercial MgO (Carlo Erba, 99%, 27 m2/g).
250 ml of distilled water were slowly added to 25 g of commercial MgO
and stirred at room temperature. The temperature was then raised to
353 K and stirring was maintained for 4 h. Excess of water was removed
by drying the sample in an oven at 358 K overnight. The resulting
Mg(OH)2 was decomposed in N2 (30 ml/min STP) to obtain high-surface
area MgO which was then treated for 18 h in N2 either at 673, 773 or 873 K
to give samples MgO-673, MgO-773 and MgO-873, respectively.

2.2 Catalyst characterization


The decomposition of Mg(OH)2 was investigated by differential thermal
analysis (DTA) using a Shimadzu DT30 analyzer, by temperature
programmed decomposition (TPDe) using a flame ionization detector
with a methanation catalyst (Ni/Kieselghur) operating at 673 K and by X-ray
diffraction (XRD) in a Shimadzu XD-D1 diffractometer equipped with
Cu-Ka radiation source (l = 0.1542 nm) and a high temperature chamber.
Samples characterized by X-ray diffraction were heated at 5 K/min until
773 K, taking diffractograms at 298, 373, 573, 673 and 773 K.

Catalysis, 2014, 26, 1–28 | 3


View Online

Surface areas and pore volumes were measured by N2 physisorption


at its boiling point using the BET method and Barret-Joyner-Halender
(BJH) calculations, respectively, in an Autosorb Quantochrome 1-C
sorptometer. The crystalline structure properties of MgO-x samples were
determined by X-ray diffraction (XRD) using the instrument described
above. Analysis was carried out using a continuous scan mode at 21/min
over a 2y range of 201–801. Scherrer equation was used to calculate the
Published on 24 February 2014 on http://pubs.rsc.org | doi:10.1039/9781782620037-00001

mean crystallite size of the samples.


CO2 adsorption site densities and binding energies were determined
from temperature-programmed desorption (TPD) of CO2 preadsorbed at
room temperature. MgO-x samples were pretreated in situ in a N2 flow
at its corresponding stabilization temperature (673, 773 or 873 K), cooled
to room temperature, and then exposed to a mixture of 3% CO2/N2 until
surface saturation was achieved (10 min). Weakly adsorbed CO2 was
removed by flushing in N2 during 1 h. Finally, the temperature was
increased to 773 K at 10 K/min. The desorbed CO2 was converted to
methane by means of a methanation catalyst (Ni/Kieselghur) operating at
673 K and monitored using a flame ionization detector.
Downloaded on 16/02/2015 11:50:03.

The chemical nature of adsorbed surface CO2 species was determined


by infrared (IR) spectroscopy after CO2 adsorption at 298 K and
sequential evacuation at increasing temperatures. Experiments were
carried out using an inverted T-shaped cell containing the sample pellet
and fitted with CaF2 windows. Data were collected in a Shimadzu FTIR
Prestige-21 spectrometer. The absorbance scales were normalized to
20-mg pellets. Each sample was pretreated in vacuum at its corres-
ponding stabilization temperature and cooled to room temperature, after
which the spectrum of the pretreated catalyst was obtained. After
admission of 5 kPa of CO2 to the cell at room temperature, the samples
were evacuated at increased temperatures, and the resulting spectrum
was recorded at room temperature. Spectra of the adsorbed species were
obtained by subtracting the catalyst spectrum.

2.3 Catalytic testing


2.3.1 Cross-aldol condensation of citral with acetone. The cross-aldol
condensation of citral (Millennium Chemicals, 95% geranial þ neral)
with acetone (Merck, p.a.) was carried out at 353 K under autogenous
pressure (E250 kPa) in a batch Parr reactor, using acetone/citral =
49 (molar ratio) and catalyst/(citral þ acetone) = 1 wt% ratio. The reactor
was assumed to be perfectly mixed and interparticle and intraparticle
diffusional limitations were verified to be negligible. Reaction products
were analyzed by gas chromatography in a Varian Star 3400 CX chro-
matograph equipped with a FID and a Carbowax Amine 30 M capillary
column. Samples of the reaction mixture were extracted every 30 min and
analyzed during the 6-h reaction. The main product of the citral/acetone
reaction was pseudoionone, PS (cis- and trans-isomers). Moreover, di-
acetone alcohol and mesityl oxide were simultaneously produced from
self-condensation of acetone. Selectivities (Sj, mol of product j/mol of
citral reacted) were calculated as Sj (%) = Cj  100/SCj, where Cj is the

4 | Catalysis, 2014, 26, 1–28


View Online

concentration of product j. Yields (Zj, mol of product j/mol of citral fed)


were calculated as Zj = SjXCit, where XCit is the citral conversion.
2.3.2 Glycerolysis of methyl oleate. The transesterification of methyl
oleate, FAME, (Fluka, W60.0%, with 86% total C18 þ C16 esters as
determined by gas chromatography) with glycerol (Aldrich, 99.0%,) was
carried out at 493 K in a seven-necked cylindrical glass reactor that
allows: separate loading of the two reactants and the catalyst, stirrer,
Published on 24 February 2014 on http://pubs.rsc.org | doi:10.1039/9781782620037-00001

thermocouple, in-out of inert gas to eliminate methanol of the gas phase,


and periodical product sampling.
Glycerol/FAME molar ratio of 4.5 and a catalyst/FAME ratio
(Wcat/n0FAME) of 30 g/mol were used. The reactor was operated in a semi-
batch regime at atmospheric pressure under N2 (35 cm3/min). Liquid
reactants were introduced into the reactor and flushed with nitrogen;
then the reactor was heated to reaction temperature under stirring
(700 rpm). Reaction products were a- and b-glyceryl monooleates (MG),
1,2- and 1,3-glyceryl dioleates (diglycerides) and glyceryl trioleate
(triglyceride). Reactant and products were analyzed by gas chroma-
Downloaded on 16/02/2015 11:50:03.

tography in a SRI 8610C gas chromatograph equipped with a flame


ionization detector, on-column injector port and a HP-1 Agilent
Technologies 15 meter  0.32 mm  0.1 mm capillary column after sily-
lation to improve compound detectability, as detailed elsewhere [49].
Twelve samples of the reaction mixture were extracted and analyzed
during the 8-h catalytic run.
2.3.3 Hydrogen transfer reduction of mesityl oxide with 2-propanol.
The gas-phase mesityl oxide/2-propanol reaction was conducted at 573 K
and atmospheric pressure in a fixed bed reactor. MgO-x samples sieved at
0.35–0.42 mm were pretreated in N2 at the corresponding calcination
temperatures for 1 h before reaction in order to remove adsorbed H2O
and CO2. The reactants, mesityl oxide (Acros 99%, isomer mixture of
mesityl oxide/isomesityl oxide = 91/9) and 2-propanol (Merck, ACS,
99.5%), were introduced together with the proper molar composition via
a syringe pump and vaporized into flowing N2 to give a N2/IPA/MO = 93.4/
6.6/1.3, kPa ratio. Reaction products were analyzed by on-line gas
chromatography in a Varian Star 3400 CX chromatograph equipped
with a flame ionization detector and a 0.2% Carbowax 1500/80–100
Carbopack C column. Main reaction products from mesityl oxide
conversion were identified as the two unsaturated alcohol isomers (UOL,
4-methyl-3-penten-2ol and 4-methyl-4-penten-2ol), isomesityl oxide,
methyl isobutyl ketone, and methyl isobutyl carbinol.

3 Results and discussion


3.1 Generation and characterization of active sites in MgO
3.1.1 Generation of active sites. The base site properties of MgO
depend on the preparation method. Usually, MgO is obtained by
decomposition of Mg(OH)2 that in turn is produced by different methods
such as chemical vapor deposition (CVD), sol-gel, precipitation, and MgO
hydration. It has been reported [50] that after Mg(OH)2 decomposition at

Catalysis, 2014, 26, 1–28 | 5


View Online

high temperature (1023 K), the relative distribution of surface low-


coordination O2 anions is shifted toward the less coordinated ions
along the series MgO-CVDoMgO-hydrationEMgO-precipitationoMgO-
sol-gel. The same order was observed for MgO activity to convert
2-methylbut-3-yn-2-ol into acetone and acetylene, a base-catalyzed
reaction [50]. The density and strength of base sites on MgO may also
be regulated by controlling both the Mg(OH)2 decomposition and MgO
Published on 24 February 2014 on http://pubs.rsc.org | doi:10.1039/9781782620037-00001

activation conditions. For example, Vidruk et al. [51] reported that


densification of Mg(OH)2 before its dehydration to obtain MgO generates
a significant increase of surface basicity. We have recently investigated
[52] the effect of calcination temperature of MgO obtained by Mg(OH)2
decomposition on its base and catalytic properties.
3.1.1.1 Thermal decomposition of Mg(OH)2. The thermal de-
composition of Mg(OH)2 precursor was studied by XRD. The diffracto-
grams in Fig. 1 showed that the Mg(OH)2 brucite structure was stable up
to about 573 K, but then, between 573 and 673 K, decomposed to MgO.
Figure 1 also shows that the MgO stabilized at 773 K during 18 h is more
Downloaded on 16/02/2015 11:50:03.

crystalline than that obtained by dynamic heating up to the same


temperature. Consistently, characterization by DTA technique showed
that the Mg(OH)2 heating exhibits an endothermic peak between 573 and
673 K arising from the solid decomposition [52]. On the other hand,
TPDe experiments revealed the presence of evolved CO2 in the 573–673 K

Mg(OH)2
MgO

stabilized at
773 K for 18 h

773 K

673 K
Intensity

573 K

298 K

30 35 40 45 50
2θ (°)

Fig. 1 XRD diffraction patterns of Mg(OH)2 decomposition.

6 | Catalysis, 2014, 26, 1–28


View Online

decomposition region, thereby suggesting that the Mg(OH)2 surface is


reversibly carbonated by interaction with atmospheric CO2. All these
results showed that the thermal treatment of Mg(OH)2 between 575 and
675 K decomposes the solid into crystalline MgO and eliminates
adsorbed carbonate species.

3.1.1.2 Physical properties of MgO samples calcined at increasing


Published on 24 February 2014 on http://pubs.rsc.org | doi:10.1039/9781782620037-00001

temperatures. The physical propertied of MgO-x samples are presented in


Table 1. The MgO surface area decreased with calcination temperature,
from 196 m2/g (MgO-673) to 169 m2/g (MgO-873), while both the mean
pore size and the pore volume increased with calcination temperature. The
XRD patterns of MgO-x samples exhibited only one crystalline species of
MgO periclase. The face-centered cubic unit cell dimensions for MgO-x
samples given in Table 1 show that the lattice parameter (a) decreased with
calcination temperature. Contraction of the MgO unit cell was accom-
panied by the increase of crystallite diameter and the sample crystallinity.
Data in Table 1 showed that, as expected, the increase of the calcination
temperature generated more ordered structures.
Downloaded on 16/02/2015 11:50:03.

3.1.2 Characterization of base sites of MgO-x samples. The surface


basic properties of MgO-x samples were probed by TPD of CO2 and by
FTIR of CO2 preadsorbed at room temperature and desorbed at
increasing temperatures. Figure 2 shows the IR spectra obtained for
MgO-x samples that reveal the presence of at least three different species:
unidentate carbonate, bidentate carbonate and bicarbonate [53–57].
Unidentate carbonate formation requires isolated surface O2 ions, i.e.,
low-coordination anions, such as those present in corners or edges and
exhibits a symmetric O-C-O stretching at 1360–1400 cm1 and an asym-
metric O-C-O stretching at 1510–1560 cm1. Bidentate carbonate forms
on Lewis acid-Brønsted base pairs (Mg2þ-O2 pair site), and shows a
symmetric O-C-O stretching at 1320–1340 cm1 and an asymmetric O-C-O
stretching at 1610–1630 cm1. Bicarbonate species formation involves
surface hydroxyl groups and shows a C-OH bending mode at 1220 cm1
as well as symmetric and asymmetric O-C-O stretching bands at
1480 cm1 and 1650 cm1, respectively. Bicarbonate was the most labile
species and disappeared on all the samples after evacuation at 373 K. In
contrast, both the unidentate and bidentate carbonates remained on the
surface after evacuation at 473 K, but only the unidentate carbonate
bands were observed upon evacuation at higher temperatures. These
results suggest the following strength order for surface basic sites: low-
coordination O2 anionsWoxygen in Mg2þ-O2 pairsWOH groups. On a
perfect MgO (1 0 0) surface, Mg2þ and O2 are five coordinated ions (Mg5c
and O5c) but on the surface of the high-surface area MgO catalysts used
here, both ions are also present with coordination numbers (L) lower
than 5 depending on the location in corners or edges. Specifically, L is 5,
4 or 3 for ions in terrace, edge or corner sites, respectively, as shown in
Fig. 3 [58, 59]. Other authors have confirmed using HRTEM that MgO
particles prepared from precipitation of Mg(OH)2 are plenty of surface
defects [27]. In Fig. 2, the overlapping adsorption bands giving rise to

Catalysis, 2014, 26, 1–28 | 7


Published on 24 February 2014 on http://pubs.rsc.org | d
Downloaded on 16/02/2015 11:50:03.
8 | Catalysis, 2014, 26, 1–28

Table 1 Physical and basic properties of MgO-x samples.

Textural characterization XRD analysis Base site density (mmol/m2)a


Surface area Pore volume Lattice Crystallite
Sample (m2/g) (ml/g) parameter, a (Å) size (Å) Crystallinity (%) Weak nOH Medium nMg-O Strong nO Total nb

MgO-673 196 0.30 4.243 74.3 85.4 0.71 1.21 2.66 4.58
MgO-773 189 0.38 4.221 76.5 86.6 0.54 1.26 1.66 3.46
MgO-873 169 0.44 4.214 143.0 93.4 0.51 1.78 1.19 3.13
a
By TPD of CO2.
View Online
O
MgO-873
C
O O
M
O O
C
O O
M C
O OH
Published on 24 February 2014 on http://pubs.rsc.org | doi:10.1039/9781782620037-00001

a
b
c
d
Absorbance (a.u.)

MgO-773

a
b
Downloaded on 16/02/2015 11:50:03.

c
d

MgO-673

b
c
d

1800 1600 1400 1200


n (cm –1)

Fig. 2 Infrared spectra of CO2 adsorbed on MgO-x catalysts upon evacuation at in-
creasing temperatures: (a) 298 K, (b) 373 K, (c) 473 K, (d) 573 K.

broad bands for both, unidentate and bidentate carbonate species, sug-
gest the presence of surface sites with different coordination numbers,
i.e., in different chemical environment, that bind CO2 with a distribution
of basic strength. In contrast, narrower and smoother bands were ob-
tained on more crystalline MgO particles resulting from calcination at
higher temperatures [27, 55, 60].
From the spectra of Fig. 2 we determined the unidentate carbonate/
bidentate carbonate band intensity ratios (U.C./B.C.); the obtained values
for MgO-673, MgO-773 and MgO-873 samples are plotted in Fig. 4. It is
observed that the (U.C./B.C.) intensity ratio on MgO calcined at 673K was
2.5 and then decreased with the calcination temperature. This result may
be interpreted by considering that the decomposition of Mg(OH)2
at relatively low temperature, i.e. 673 K, generates hydroxylated MgO

Catalysis, 2014, 26, 1–28 | 9


Published on 24 February 2014 on http://pubs.rsc.org | doi:10.1039/9781782620037-00001 View Online
Downloaded on 16/02/2015 11:50:03.

Fig. 3 Scheme of a stepped MgO (1 0 0) surface with OLc and MgLc ions in different
positions (L: coordination number). Terrace sites: O5c, Mg5c; edge sites: O4c, Mg4c; corner
sites: O3c, Mg3c.

2.8
2.6
2.4
nO/nMg-O (by TPD of CO2)
U.C./B.C. (by IR of CO2)

2.4

2.2 2.0

2.0 1.6

1.8 1.2
1.6
0.8
1.4
0.4
650 700 750 800 850 900
Temperature (K)

Fig. 4 Strong/medium-strength base site and U.C./B.C. ratios as a function of calcination


temperature. (B.C.: bidentate carbonate; U.C.: unidentate carbonate).

containing a high concentration of low-coordination O2 sites located


on defects of the crystalline solid surface. Then, the increase of the
calcination temperature would remove OH groups and also surface solid
defects creating a smoother and thermodynamically more stable struc-
ture, as suggested by the XRD data of Table 1. This interpretation is
consistent with the results reported by Morterra et al. [54] on MgAl2O4
and by Evans and Whateley [61] on MgO. These authors investigated by
IR of CO2 the role of surface hydroxylation on the generation of strong
basic sites and concluded that the strong basicity, responsible for
unidentate carbonate formation, is promoted by the presence of surface

10 | Catalysis, 2014, 26, 1–28


View Online

5
MgO-873

CO2 desorption rate (μmol/hm2)


MgO-773
Published on 24 February 2014 on http://pubs.rsc.org | doi:10.1039/9781782620037-00001

MgO-673

strong

medium
weak
Downloaded on 16/02/2015 11:50:03.

300 400 500 600 700


Temperature (K)

Fig. 5 TPD profiles of CO2 on MgO-x samples. CO2 adsorption at 298 K, 10 K/min
heating rate.

OH groups. Then, the observed loss of unidentate carbonate formation


centers (low coordination surface O2 ions) when the calcination
temperature is increased can be ascribed to both the elimination of
surface defects and the enhancement of surface dehydroxylation.
A measure of the number and strength distribution of basic sites
on MgO-x samples was obtained by TPD of CO2 preadsorbed at room
temperature. The CO2 desorption rate as a function of desorption
temperature is presented in Fig. 5. The total base site densities of
desorbed CO2 (nb, mmol/m2) were measured by integration of TPD curves
in Fig. 5 and are reported in Table 1. It is observed that nb decreased
with calcination temperature, from 4.58 mmol/m2 (MgO-673) to
3.13 mmol/m2 (MgO-873), thereby confirming a solid surface transfor-
mation that goes beyond the mere coalescence of the pore structure.
Based on the previous IR characterization data, the TPD profiles of
Fig. 5 were deconvoluted in three desorption peaks: a low temperature
peak at 390 K, assigned to bicarbonates formed on surface OH groups, a
middle-temperature peak at 440 K attributed to bidentate carbonates
desorbed from Mg2þ-O2 pairs, and a high-temperature peak at 550 K
resulting from unidentate carbonates released from low-coordination
O2 anions. By integrating these three CO2 TPD peaks we determined the
density of strongly basic low coordination (O3c and O4c) anions identified
as nO in Table 1, medium strength Mg5c-O5c pair sites, nMg-O, and weak
OH groups, nOH. Results in Table 1 show that nO and nOH decreased while
nMg-O increased with the MgO calcination temperature. In Fig. 4 we plotted
the nO/nMg-O ratio as a function of calcination temperature; it is observed

Catalysis, 2014, 26, 1–28 | 11


View Online

that nO/nMg-O values decreased with the calcination temperature following


a similar trend that the (U.C./B.C.) ratio determined by IR spectroscopy.
In summary, all these results show that the decomposition of Mg(OH)2
at 673 K generates hydroxylated MgO containing predominantly high-
strength low-coordination O2 basic sites located on defects of the crys-
talline solid surface. The increase of the calcination temperature up to
873 K removes OH groups and also surface solid defects creating more
Published on 24 February 2014 on http://pubs.rsc.org | doi:10.1039/9781782620037-00001

stable structures that contain a higher concentration of medium-strength


Mg2þ-O2 basic pair sites. Thus, the density, nature and strength of MgO
surface basic sites may be regulated by modifying the solid calcination
temperature.
Finally, it is significant to note here that in a previous work we have
characterized the acid properties of MgO-673 by NH3 TPD and FTIR of
adsorbed pyridine [62]. We observed that MgO-673 contained only weak
Lewis Mgþ2 acid sites; the density of Mg2þ sites as determined by NH3
TPD was 0.14 mmol/m2, i.e. about 30 times lower than the density of base
sites determined by CO2 TPD (Table 1, nb = 4.58 mmol/m2).
Downloaded on 16/02/2015 11:50:03.

3.2 Catalytic results on MgO-x catalysts


In order to investigate the effect of MgO calcination temperature on
catalyst activity, we carried out two base-catalyzed reactions on our MgO-x
samples: the liquid-phase aldol condensation of citral with acetone
and the gas-phase hydrogen transfer reduction of mesityl oxide with 2-
propanol. For both reactions, catalysts were treated at their calcination
temperatures prior to performing the catalytic tests.
3.2.1 Cross-aldol condensation of citral with acetone. The aldol
condensation of citral with acetone produces pseudoionone (Fig. 6), a
valuable acyclic intermediate for the synthesis of ionones which are
extensively used as pharmaceuticals and fragrances [63]. The reaction
was commercially carried out using diluted bases such as NaOH, Ba(OH)2
or LiOH [64, 65], but it is also efficiently catalyzed on solid bases [13, 66–68].
Here, the liquid-phase citral/acetone reaction was performed on the
MgO-x samples of Table 1. Figure 7 shows the evolution of pseudoionone
yields (ZPS) as a function of reaction time. At the end of the 6-h catalytic
tests, citral conversion was 96%, 80% and 75% for MgO-673, MgO-773
and MgO-873 samples, respectively (Table 2). From the curves of Fig. 7,
we determined the initial pseudoionone formation rate (r0PS, mol/h m2)
through the initial slopes according to:
 
n0Cit dZPS
r0PS =
Wcat Sg dt t¼0

Fig. 6 Synthesis of pseudoionone by citral/acetone aldol condensation.

12 | Catalysis, 2014, 26, 1–28


View Online
100

80

Pseudoionone Yield (%)


60
Published on 24 February 2014 on http://pubs.rsc.org | doi:10.1039/9781782620037-00001

40

MgO-673
20 MgO-773
MgO-873

0
0 1 2 3 4 5 6
Time (h)

Fig. 7 Aldol condensation of citral with acetone: pseudoionone yield as a function of time
(T = 353 K, n0Acet = 0.8 moles, n0Cit = 0.016 moles, WCat. = 0.5 g).
Downloaded on 16/02/2015 11:50:03.

where Wcat is the catalyst weight and n0Cit are the initial moles of citral.
The obtained r0PS values (Table 2) decreased with the calcination
temperature, following a trend similar to the density of strong base sites
shown in Table 1 (nO values). In all the cases, the initial selectivities to
pseudoionones were about 100% showing that the conversion of citral
via other reactions than its condensation with acetone is negligible. The
observed proportionality between r0PS and nO suggests that under initial
conditions the rate-determining step for the citral/acetone reaction
toward pseudoionones is promoted by strongly O2 basic sites, which is
in agreement with the results reported elsewhere on MgO-based cata-
lysts [13, 62]. The function of surface O2 sites is to abstract the a-proton
from acetone, forming a carbanion that consecutively attacks the car-
bonyl group of the contiguously adsorbed citral molecule, as depicted in
Fig. 8. Then a b-hydroxyl ketone intermediate is expected to form; how-
ever, this compound was never observed among the reaction products
under the reaction conditions of this work. Therefore, this unstable
intermediate is assumed to rapidly dehydrate, forming pseudoionone
and water and regenerating the active sites on the catalyst surface. The
role of surface Mg2þ sites is to provide adsorption sites for acetone
through its carbonyl group and to stabilize the reaction intermediates
(Fig. 8).

3.2.2 Hydrogen transfer reduction of mesityl oxide with 2-propanol.


The selective synthesis of secondary unsaturated (UOL) alcohols
from reduction of alkyl vinyl ketones is an important process for
pharmaceutical, fragrance and polymer industries. This reaction is
hardly achieved on noble metals by conventional hydrogenation that
uses high-pressure H2 in multiphase batch reactors because reduction of
the C¼C bond is thermodynamically and kinetically favored over that of
the C¼O group [69, 70]. The substituent at the carbonyl hinders

Catalysis, 2014, 26, 1–28 | 13


Published on 24 February 2014 on http://pubs.rsc.org | d
Downloaded on 16/02/2015 11:50:03.
14 | Catalysis, 2014, 26, 1–28

Table 2 Catalytic activity data on MgO-x samples.

Reaction
Citral/acetone (liquid phase)a MO/2-propanol (gas phase)b FAME/glycerol (liquid phase)c
Initial PS formation rate, r0PS UOL formation rate, rUOL, Initial MG formation rate
d 2 2 e
Catalyst XCit, % mmol/h g mmol/h m mmol/h g mmol/h m XFAME (%) mmol/h g mmol/h m2

MgO-673 96 93.5 0.477 18.2 0.093 93 28.6 0.146


MgO-773 80 76.2 0.403 18.3 0.097 84 21.2 0.112
MgO-873 75 62.9 0.372 18.6 0.110 76 15.2 0.090
a
T = 353 K, n0DMK = 0.8 moles, n0Cit = 0.016 moles, WCat. = 0.5 g.
b
T = 573 K, P = 101.3 kPa, N2/IPA/MO = 93.4/6.6./1.3, W=F0MO = 15 g h/mol.
c
T = 493 K; Gly/FAME = 4.5; Wcat/n0FAME = 30 g/mol K.
d
At the end of the 6-h catalytic runs.
e
After 3 h of reaction.
Published on 24 February 2014 on http://pubs.rsc.org | doi:10.1039/9781782620037-00001 View Online

Fig. 8 Reaction mechanism for citral/acetone aldol condensation.

O OH OH O
H 3C H3 C
C=CH-C-CH3 + H3C-CH-CH3 C=CH-C-CH3 + H3C-C-CH3
H 3C H3C
4-methyl-3-penten-2-one 2-propanol 4-methyl-3-penten-2 ol Acetone
(MO) (UOL)

Fig. 9 Unsaturated alcohol (UOL) synthesis by hydrogen transfer reduction (HTR) of


Downloaded on 16/02/2015 11:50:03.

mesityl oxide (MO) with 2-propanol.

coordination of the C¼O bond on the surface thereby decreasing


the chemoselectivity for the C¼O bond saturation [71, 72]. In addition,
the consecutive UOL isomerization to the corresponding saturated
ketone is usually an unavoidable side reaction on metallic catalysts [73].
Hydrogen transfer reduction (HTR) reactions is an alternative route
for the catalytic synthesis of UOL by asymmetric reduction of the
corresponding a ketone. In the HTR reaction, the carbonyl compound
(oxidant) is contacted with a hydrogen donor (reductant) at mild
conditions in liquid or gas phase without supply of molecular hydrogen.
Heterogeneously catalyzed HTR of unsaturated carbonyl compounds
would occur on metal oxides via a Meerwein-Ponndorf-Verley mech-
anism, which involves the selective reduction of the C¼O bond
preserving the C¼C bond [74, 75]. In particular, we have studied
the gas-phase HTR of 2-cyclohexenone and mesityl oxide (MO) with
2-propanol toward the corresponding unsaturated alcohol on base,
acid-base and metal/acid-base catalysts [18, 76, 77]. Here, we present
the results obtained for the gas-phase mesityl oxide/2-propanol reaction
(Fig. 9) on the MgO-x samples of Table 1.
When both reactants (MO and 2-propanol) are co-fed to the reactor, in
addition to the reaction of Fig. 9 several parallel or consecutive reactions
can take place, such as:
i) MO double bond isomerization

Catalysis, 2014, 26, 1–28 | 15


View Online

ii) Selective C¼C bond reduction of MO or i-MO to MIBK

iii) Simultaneous C¼C and C¼O bond reduction of MO or i-MO to


Published on 24 February 2014 on http://pubs.rsc.org | doi:10.1039/9781782620037-00001

MIBC:

iv) Aldol condensation reactions between C6 carbonyl compounds and


acetone toward C9 compounds.
To obtain insight on the reaction pathways of the MO/2-propanol
reaction, we investigated the effect of contact time (W/F0MO) on product
Downloaded on 16/02/2015 11:50:03.

distribution over MgO-773 at 523 K by varying W/F0MO between 2.0 and


42.0 gcat h/mol MO [77]. In Fig. 10 we plotted the yields for MO reactions
as a function of contact time. UOL formed fast and directly from MO and
i-MO, and the UOL yield increased with W/F0MO up to about 42% without
reaching any maximum, thereby suggesting that UOL does not partici-
pate significantly in consecutive reactions on MgO-773. MIBK yields of
less than 1% were measured regardless of the conversion level thereby
confirming that selective reduction of the C¼C bond of mesityl oxide is
unlikely on MgO. The MIBC curve in Fig. 10 is consistent with direct
formation from MO at low conversions but also from i-MO or to a lesser
extend from UOL at high contact times. Finally, the initial zero slope for
C9 product formation reveals that aldol condensation compounds are

UOL
40

30 MIBC
Yields (ηi, %)

20
i-MO

10
MIBK C9

0
0 10 20 30 40
W/F0MO (gcat h/mol MO)

Fig. 10 HTR of of mesityl oxide with 2-propanol: Yields as a function of contact time
(523 K, 100 kPa, N2/2P/MO = 93.4/6.6/1.3).

16 | Catalysis, 2014, 26, 1–28


Published on 24 February 2014 on http://pubs.rsc.org | doi:10.1039/9781782620037-00001 View Online

Fig. 11 Scheme of the reaction pathways on MgO for the HTR of mesityl oxide (MO).

Fig. 12 UOL formation by MPV mechanism.


Downloaded on 16/02/2015 11:50:03.

formed in consecutive pathways. Figure 11 presents a simplified reaction


network for the HTR of mesityl oxide with 2-propanol on MgO that is
consistent with the catalytic results showed in Fig. 10.
With the aim of establishing the effect of the calcination temperature
of MgO on the catalyst activity for gas-phase HTR reactions we performed
the MO/2-propanol reaction on our MgO-x samples using a contact time
of 15 g h/mol. The obtained UOL formation rates (rUOL) are shown in
Table 2. It is observed that rUOL, expressed either in mass or area basis,
slightly increased with the calcination temperature, similarly to the
evolution of medium-strength basic sites shown in Table 1 (nMg-O values).
This result suggests that the rate limiting step for the formation of UOL is
promoted on Mg2þ-O2 pair sites. In fact, as it is shown in Fig. 12 the
Mg2þ-O2 pair sites would promote formation of the six-atom cyclic
intermediate needed in Meerwein-Ponndorf-Verley mechanism for pref-
erentially transferring hydrogen from the 2-propanol donor molecule to
the C¼O bond of mesityl oxide. Mesityl oxide adsorbs via the C¼O bond
on a weak Lewis acid Mg2þ cation, whereas 2-propanol adsorbs non-
dissociatively on a vicinal Mg2þ-O2 pair, giving rise to the required
bimolecular six-atom cyclic intermediate [77]. Then, hydride transfer
occurs without participation of surface H fragments, selectively forming
the unsaturated alcohol.
In summary, our results above show that the density, nature and
strength of MgO surface basic sites may be regulated by modifying the
solid calcination temperature. But the effect of calcination temperature on
the MgO catalytic properties depends on the basicity requirements for the
rate-limiting step of the base-catalyzed reaction. For example, the activity
for the liquid-phase synthesis of pseudoionones by condensation of citral
with acetone diminishes with MgO calcination temperature because this
reaction is predominantly promoted on strongly basic O2 sites. In con-
trast, the synthesis of 4-methyl-3-penten-2ol by the gas-phase hydrogen
transfer reduction of mesityl oxide with 2-propanol is improved by

Catalysis, 2014, 26, 1–28 | 17


View Online

increasing the MgO calcination temperature because the reaction inter-


mediate is formed on medium-strength Mg2þ-O2 pair basic sites.

3.3 DFT molecular modeling studies of MgO active sites


We performed DFT calculations to obtain additional information on the
role played by the MgO active sites on the kinetics of base-catalyzed
reactions. Specifically, we present molecular modeling studies on our
Published on 24 February 2014 on http://pubs.rsc.org | doi:10.1039/9781782620037-00001

MgO-x catalysts for the synthesis of monoglycerides (MG) from glycerol


(Gly) by transesterification (glycerolysis) of methyl oleate (C18 : 1), an
unsaturated fatty acid methyl ester (FAME) (Fig. 13).
Monoglycerides present surfactant and emulsifying properties that
help hydrophilic and lipophilic substances mix together. Therefore, they
can be used in food, detergent, plasticizer, cosmetic and pharmaceutical
formulations [78]. The commercial liquid-catalyzed synthesis route to
produce MG involves strong mineral bases such as Ca(OH)2 and KOH;
this process yields only 40–60% MG, the rest being diglycerides and tri-
glycerides, and entails concerns related to corrosion and disposal of
spent base materials. The use of solid catalysts for MG synthesis presents
Downloaded on 16/02/2015 11:50:03.

not only the known environmental and practical advantages but also
provides the opportunity to increase the MG yield. However, industrial
implementation of heterogeneously catalyzed processes for FAME
glycerolysis able to efficiently replace the use of liquid bases is still a
challenge. Previous works have discussed the different routes for MG
synthesis by esterification of fatty acids or by transesterification of
triglycerides or fatty acid methyl esters [7, 79]. The base-catalyzed MG
synthesis from Gly using FAME instead of fatty acids or triglycerides has
several advantages, e.g., FAME is less corrosive than FA, has lower
hydrophobic character than triglycerides, and exhibits higher miscibility
with glycerol; therefore, the process can be carried out at lower tem-
peratures than TG transesterification. Furthermore, the reaction route
from FAME yields MG with a definite acyl group composition (FAME are
easier to separate by fractional distillation than fatty acids) whereas in TG
glycerolysis the products contain the acyl group distribution of the oil or
fat [80]. MgO-based catalysts such as Mg/MCM-41 and Mg-Al mixed oxi-
des have been investigated for the MG synthesis from glycerolysis of
FAME [79, 81]. In particular, we have studied the glycerolysis of methyl
oleate on MgO-based catalysts [7, 49, 82] and reported the reaction con-
ditions needed to implement this reaction in a four-phase reactor under
kinetic control and to reach maximum MG yields. Here, we present the
results obtained on our MgO-x samples to get insight into the base site
strength requirements for glycerolysis reactions.

Fig. 13 Synthesis of monoglycerides by transesterification of FAME with glycerol.

18 | Catalysis, 2014, 26, 1–28


View Online

3.3.1 FAME glycerolysis: Catalytic results on MgO-x catalysts. Figure


14 shows FAME and Gly conversions and yields on MgO-773 at 493 K and
typically illustrates the time-on-stream behavior of the catalyst during the
reaction. Results show that monoglycerides can be obtained in high
yields (70%) in 8 h using MgO. Only MG was initially formed but as the
reaction proceeded, a second transesterification took place forming
diglycerides from MG and FAME. MG was obtained in higher selectivity
Published on 24 February 2014 on http://pubs.rsc.org | doi:10.1039/9781782620037-00001

than diglycerides under the present conditions, but glyceride selectivity


can be controlled by changing the Gly/FAME ratio so that to modify
the availability of both reactants in the reaction zone, as previously
discussed [7].
Similarly to Fig. 14 for MgO-773, we plotted the evolution of FAME and
Gly conversions and yields for MgO-673 and MgO-873 samples (plots not
shown here). We observed that the MgO activity decreased with calcin-
ation temperature; FAME conversion after 3 h of reaction was, in fact,
93%, 84% and 76% for samples MgO-673, MgO-773 and MgO-873, re-
spectively (Table 2). The initial MG formation rates (r0MG) were deter-
mined from the slopes at t = 0 of MG yield versus time curves and the
Downloaded on 16/02/2015 11:50:03.

results are included in Table 2. The r0MG values of Table 2 were plotted in
Fig. 15 as a function the strong base site density (nO values in Table 1).
This result suggested that under initial conditions the rate-determining
step for MG formation is essentially promoted on strong base sites,
present in corners or edges of the non-uniform surface of MgO catalysts.
To obtain more insight on the role played by the MgO surface sites in
the reaction kinetics we carried out a molecular modeling of Gly and
FAME adsorptions on MgO [83] using a cluster model that represents the
MgO surface with four different adsorption sites as depicted in Fig. 16:
the terrace site contains the Mg5c-O5c pairs (L = 5) that model the MgO
medium strength base sites; the edge and the O-apical corner sites

80
100
XFAME
FAME conversion, XFAME (%)

Yα-MG
60 80
Glyceride Yield, Yj (%)

60
40

40
Y1,3-DG
20
Yβ-MG 20

Y1,2-DG
0 0
0 1 2 3 4 5 6 7 8
Time (h)

Fig. 14 FAME conversion and glyceride yields [MgO-773, Gly/FAME = 2; T = 493 K;


Wcat/n0FAME = 11 g/mol FAME].

Catalysis, 2014, 26, 1–28 | 19


View Online

0.150
MgO-673

r 0MG (mmol/h m2)


0.125 MgO-773

0.100 MgO-873
Published on 24 February 2014 on http://pubs.rsc.org | doi:10.1039/9781782620037-00001

0.075

0.050
0.5 1.0 1.5 2.0 2.5 3.0
nO, density of strong base sites (μmol/m2)

Fig. 15 Initial monoglyceride conversion rate as a function of strong base site density.
Reaction conditions as in Fig. 14.
Downloaded on 16/02/2015 11:50:03.

Fig. 16 Clusters used for modeling the MgO (1 0 0) surface. (a) Perfect terrace site,
Mg25O25(Mg-ECP)25; (b) defective edge site, Mg22O22(Mg-ECP)19; (c) defective O-apical
corner site, Mg22O22(Mg-ECP)12; (d) defective Mg-apical corner site, Mg23O23(Mg-ECP)14.

represent the strongly basic O4c (L = 4) and O3c (L = 3) sites, respectively,


and a Mg-apical corner (Mg3c; L = 3) that models a Lewis acid site.

3.3.2 Computational details. DFT molecular orbital calculations


were carried out using the gradient corrected Becke’s three parameters
hybrid exchange functional in combination with the correlation
functional of Lee, Yang and Parr (B3LYP) [84]. The terrace site at the MgO
(100) surface was represented by the Mg25O25 (Mg-ECP)25 cluster consisting
of two layers (first layer: Mg9O16; second layer: Mg16O9). For the topological
defects at edges and corners similar modeling was used; a Mg22O22(Mg-
ECP)19 cluster was used for modeling the edge topological defect of
MgO due to the intersection of two [100] and [010] oriented planes; a
Mg22O22(Mg-ECP)12 cluster was used for modeling the oxygen corner
topological defect of MgO due to the intersection of three [100], [010]
and [001] oriented planes whereas the a similarly generated Mg23O23(Mg-
ECP)14 cluster was used for modeling a magnesium apical corner.
To take into account the Madelung field due to the rest of the extended
surface, the cluster was embedded in an array of  2 point charges. This
embedding technique was used previously for the study of both bulk and

20 | Catalysis, 2014, 26, 1–28


View Online

surface properties giving results which are in good agreement with those
obtained by periodic calculations [85, 86]. Moreover, the positive point
charges at the interface were replaced by effective core potentials (ECP)
corresponding to Mg2þ to account for the finite size of the cations and to
avoid spurious charge polarization.
The O atoms of the MgO surface that interact directly with the glycerol
or FAME molecule (all from the first layer) were described with the basis
Published on 24 February 2014 on http://pubs.rsc.org | doi:10.1039/9781782620037-00001

set 6-31 þ G(d) and Mg atoms with 6-31G(d). For the rest of the oxide
atoms in the cluster the basis set 6-31G was used. The 6-31G (d,p) basis
set was used for the molecular orbitals of Gly and FAME. The adsorption
energy of Gly or FAME (Eads) was evaluated according to the following
total energy difference: Eads = E(molecule-MgO cluster)  E(MgO cluster)  E(molecule);
where ‘‘molecule’’ is either Gly or FAME. Negative values indicate exo-
thermic adsorption.
On the other hand, the atomic net charges (q) were calculated
following the natural bond orbital (NBO) scheme [87], which gives real-
istic values for the charge partitioning. For all the systems the total
charge was zero. Also, Dq(atom) was defined for an atom of Gly or FAME as
Downloaded on 16/02/2015 11:50:03.

the atom charge difference between adsorbed and free molecule states.
All the calculations were performed using the Gaussian-03 program
package.

3.3.3 Glycerol adsorption on MgO. First-principles density-functional


calculations were performed for the free glycerol molecule and for the
adsorption of glycerol on representative terrace, edge, and Mg- and O-
corner sites of MgO. The DFT calculations for the optimized geometrical
structure of the free Gly molecule resulted in the following
intramolecular interatomic distances (d): C-C (d(C–C)E1.53 Å), C-H
(d(C–H)E1.10 Å), C-O (d(C-O)E1.42 Å) and O-H (d(O-H)E0.97 Å). For the Gly
adsorption on MgO, different initial geometries of the glycerol molecule
were evaluated depending on the orientation of the hydroxyl groups
toward the MgO surface. Results presented in Table 3 show the optimized
geometrical structures obtained for Gly adsorption through one, two or

Table 3 Adsorption energies (Eads) and bond distances (d) for Gly adsorption on terrace,
edge and O-apical corner sites of MgO (100).a

Final structure
Entry Cluster nOH(m) n m Eads (eV) dðHOs Þ (Å) dðOMgs Þ (Å) d(O–H) (Å)

1 Terrace (L = 5) 1OH(1) 1 1  0.65 1.594 3.228 1.008


2 Terrace (L = 5) 2OH(1,2) 2 1,2  0.92 1.588 2.252 1.019
3 Terrace (L = 5) 3OH 3 1,2,3  1.48 1.785 2.293 0.996
4 Edge (L = 4) 1OH(2) 1 2  1.63 1.050 1.988 1.478
5 Edge (L = 4) 2OH(1,2) 2 1,2  1.85 1.035 2.082 1.494
6 Edge (L = 4) 3OH 3 1,2,3  1.62 1.588 2.104 1.013
7 O-corner (L = 3) 1OH(1) 1 1  0.89 1.519 2.802 1.045
8 O-corner (L = 3) 2OH(1,2) 2 1,2  1.55 1.524 2.127 1.037
9 O-corner (L = 3) 3OH 3 1,2,3  0.85 0.985 2.187 1.676
a
Cluster sites as in Fig. 16; n and m: number and position of OH groups interacting with the
surface, respectively.

Catalysis, 2014, 26, 1–28 | 21


Published on 24 February 2014 on http://pubs.rsc.org | doi:10.1039/9781782620037-00001 View Online

Fig. 17 Scheme representing the formation of surface glyceroxide species from glycerol
and the FAME surface activation.

three hydroxyl groups. Species are identified as nOH(m), where n and m


Downloaded on 16/02/2015 11:50:03.

represent, respectively, the amount and position of the OH groups par-


ticipating in the surface species. Thus, m is 1 or 3 for primary hydroxyls
and 2 for the secondary one. The Eads values, the intramolecular and the
Gly-MgO distances (d) were calculated at equilibrium. The Gly/MgO
distances are depicted in the scheme of Fig. 17 as dðOMgs Þ and dðHOs Þ ,
considering the closest Gly hydroxyl interacting with the MgO surface.
Table 3 presents the results obtained on a perfect terrace site of MgO
(100) where Mg2þ and O2 ions are five-fold coordinated (L = 5). Results
show that the dðOMgs Þ bond distance diminishes with the number of
hydroxyl groups (n) interacting with the surface, probably reflecting a
higher electrostatic interaction between Gly and the MgO surface that
evidences the hydrophilic properties of MgO. Shortening of dðOMgs Þ
suggests the presence of a more stable surface species, as indicated by
the larger Eads value.
On the other hand, the values of the intramolecular O-H distance, d(O–H),
in Table 3 suggest that regardless of the adsorption species structure, the
OH groups maintained their integrity which indicates that glycerol ad-
sorbs non-dissociatively on surface terrace sites. The glycerolysis reaction
requires the rupture of an O-H bond at the Gly molecule to proceed,
as depicted in Fig. 17. Then, data in Table 3 showing that Gly is non-
dissociatively adsorbed on MgO terrace sites strongly suggest that the
Mg5c-O5c pairs do not promote the Gly/FAME reaction.
Table 3 also presents the results obtained for the glycerol adsorption
on an edge site that models the low coordination (L = 4) base sites of
MgO. The Eads values obtained for the edge site were higher than those
determined for the terrace site, thereby indicating a stronger interaction
between Gly and the surface edge site. As a result of the stronger Gly-edge
interaction, dissociative adsorption occurred for some of the postulated
geometries such as 2OH(1,2)a (Fig. 18(b), entry 5 in Table 3), where the
d(O–H) distance was longer than in free glycerol due to O-H bond breaking
with formation of both, a new surface OH between the abstracted H and a

22 | Catalysis, 2014, 26, 1–28


Published on 24 February 2014 on http://pubs.rsc.org | doi:10.1039/9781782620037-00001 View Online

Fig. 18 Optimized geometrical structures of species 2OH(1,2) of Gly adsorbed on a MgO


(100) surface: (a) non-dissociative adsorption on terrace sites; (b) dissociative adsorption
on defective edge sites; (c) non-dissociative adsorption on a defective O-apical corner site.

cluster oxygen, Os, and a surface glyceroxide on a cluster cation, Mgs


(Fig. 17). Thus, the dðOMgs Þ and dðHOs Þ distances for species 2OH(1,2)
considerably shortened compared to 1OH(2) and 3OH surface structures
of Table 3 for which non-dissociative adsorption took place.
Downloaded on 16/02/2015 11:50:03.

Glycerol adsorption was studied also on an O-apical corner (Table 3).


Comparison between the Gly adsorption through the primary OH,
structure 1OH(1), on the terrace (entry 1) and on the O-corner (entry 7)
sites, indicated that a more stable species was obtained on the O-corner
with shortening of the dðOMgs Þ and dðHOs Þ distances. However, it seems
that for the other two structures of Table 3, the Gly molecule tended to
rotate and interacted to a greater extent with the Mg5c-O5c sites of the
cluster than with O3c at the cluster corner, probably due to a steric effect
that hampered the molecule arrangement on the oxygen corner. In this
regard, Fig. 18 illustrates the adsorption of structure 2OH(1,2) of Gly on
the terrace, edge and O-corner sites of MgO (100). The increase of the
surface oxygen unsaturation from the terrace to the edge site would cause
the Gly molecule O-H bond dissociation forming much more stable
species. However, a further oxygen unsaturation increase from the edge
to the O-corner site, would give rise to a non-dissociated 2OH(1,2) species
with an Eads value in between those of the terrace and edge sites.
Consistently, the d(O–H) distances of adsorbed Gly structures on an
O-apical corner were similar to those of free Gly, suggesting that no O-H
bond breaking took place.

3.3.4 FAME adsorption on MgO. DFT calculations were also


carried out for FAME adsorption on representative terrace, edge, and
Mg- and O-corner sites of MgO. The FAME molecule used in the catalytic
experiments was methyl oleate that contains eighteen carbon atoms and
one unsaturation (C18:1). For modeling purposes, a shorter molecule
containing just five carbon atoms in the acyl chain was used (C5:0). The
optimized geometrical structures of free methyl oleate as well as of
the FAME used in the calculations are shown in Fig. 19. In the free short
C5:0 FAME molecule the calculated intramolecular interatomic
distances (d) were: C¼O (d(C¼O) = 1.212 Å), C–O (dðCOCH3 Þ ¼ 1:355 Å and
dðOCH3 Þ ¼ 1:436 Å), C–H (d(C–H)E1.09 Å) and C–C (d(C–H)E 1.53 Å).

Catalysis, 2014, 26, 1–28 | 23


Published on 24 February 2014 on http://pubs.rsc.org | doi:10.1039/9781782620037-00001 View Online
Downloaded on 16/02/2015 11:50:03.

Fig. 19 Optimized geometries for free FAME molecules. (a) FAME used in catalytic
experiments, methyl oleate (C18:1); (b) FAME used in theoretical calculations (C5:0).

Fig. 20 Optimized geometrical structures of FAME (C5:0) adsorbed on different MgO


surface sites. (a) Terrace site; (b) edge site; (c) O-apical corner site; (d) Mg-apical corner site.

All these distances were very similar to those calculated for the free
methyl oleate (C18:1) molecule.
Optimized geometrical structures for adsorption of a C5:0 FAME
molecule on terrace, edge and corner sites of MgO (100) are shown in
Fig. 20 and the results are given in Table 4. Adsorption energy values,
intramolecular bond distances and FAME-surface distances to the closest
surface atom were calculated at equilibrium. The FAME molecule is
expected to adsorb via the C¼O bond on surface Mg Lewis acid sites as
illustrated in Fig. 17. Regardless of the cluster geometry low Eads values for
FAME adsorption on MgO were obtained (Table 4), thereby indicating that
the FAME-surface interaction is weak. The Eads values in Table 4 were in fact
similar to those reported for the adsorption of non-polar low interacting
molecules such as methane and benzene on a terrace site of MgO [88].
The C¼O bond distance, d(C¼O), in adsorbed FAME molecules on ter-
race, edge and corner sites increased as the coordination number of the

24 | Catalysis, 2014, 26, 1–28


View Online

Table 4 Adsorption energies (Eads), carbonyl oxygen charge difference DqðOC¼O Þ and
bond distances (d) for FAME adsorption on terrace, edge and Mg- and O-apical corner
sites of MgO (100).a

Entry Cluster site Eads (eV) DqðOC¼O Þ (a.u.) dðOC¼O Mgs Þ (Å) d(C¼O) (Å)

1 Terrace (L = 5)  0.05  0.02 2.474 1.219


2 Edge (L = 4)  0.02  0.04 2.298 1.224
3 O-corner (L = 3)  0.56  0.07 2.189 1.227
Published on 24 February 2014 on http://pubs.rsc.org | doi:10.1039/9781782620037-00001

4 Mg-corner (L = 3)  0.67  0.07 2.148 1.223


a
Cluster sites as in Fig. 16.

ions in the cluster decreased from L = 5 to L = 3, suggesting a stronger


FAME-surface interaction. Consistently, the O–Mgs bond distance,
dðOC¼O Mgs Þ , decreased when FAME was adsorbed on lower coordination
surface ions. Nevertheless, in all the cases the d(C¼O) distances remained
close to that of the free molecule (1.212 Å) which shows that the integrity
of the adsorbed FAME molecule is preserved. In line with these results,
the qðOC¼O Þ values in Table 4 indicated that the oxygen of the C¼O bond
Downloaded on 16/02/2015 11:50:03.

gained some negative charge as a consequence of the adsorption process.


However, in all the cases low qðOC¼O Þ values were obtained thereby sug-
gesting that polarization does not proceed to a significant extent on any
cluster geometry.
In summary, DFT calculations predict that the proton abstraction
from the glycerol hydroxyl groups required in the glycerolysis reaction
(Fig. 16) would preferentially occur on low coordination O2 (strong base
O4c sites located on edges), in agreement with the catalytic results
presented in Fig. 15 and Table 2. FAME adsorption on MgO is weak, even
on low coordination Mg3c and O3c sites. Therefore, the Gly/FAME reaction
would proceed through a mechanism in which the most relevant
adsorption step is that of glycerol.

4 Conclusions
The density, nature and strength of surface basic sites on MgO obtained
from Mg(OH)2 decomposition may be regulated by modifying the solid
calcination temperature. Decomposition of Mg(OH)2 at 673 K generates
hydroxylated MgO containing mainly strong O2 basic sites located in
surface defects such as corners and edges of the crystalline solid surface.
The increase of the calcination temperature removes OH groups and also
surface solid defects creating more stable structures that contain a higher
concentration of medium-strength Mg2þ-O2 basic pair sites.
The effect of calcination temperature on the MgO activity and
selectivity for a given base-catalyzed reaction depends on the basicity
requirements for the rate-limiting step of the reaction mechanism. For
example, the activity for the liquid-phase synthesis of monoglycerides by
glycerolysis of methyl oleate as well as that of pseudoionones by
condensation of citral with acetone diminish with MgO calcination
temperature because both reactions occur predominantly on strong basic
O2 sites. In contrast, the gas-phase hydrogen transfer reduction of

Catalysis, 2014, 26, 1–28 | 25


View Online

mesityl oxide with 2-propanol is improved by increasing the calcination


temperature because formation of the six-atom cyclic intermediate
needed in Meerwein-Ponndorf-Verley mechanism for transferring
hydrogen from the 2-propanol donor molecule to the C¼O bond of
mesityl oxide is promoted on medium-strength Mg2þ-O2 pair sites.
First-principles density-functional calculations were carried out to
obtain more insigth on the role played by the MgO active sites for the
Published on 24 February 2014 on http://pubs.rsc.org | doi:10.1039/9781782620037-00001

glycerolysis of methyl oleate (FAME). DFT calculations were performed


for the adsorption of glycerol and FAME on representative terrace, edge,
and Mg- and O-corner sites of MgO. In agreement with catalytic results,
calculations predict that dissociative chemisorption of glycerol with O–H
bond breaking occurs only on low coordination O2 surface sites (edge
sites) whereas nondissociative adsorption takes place on medium-
strength base sites such as those of terrace sites. The DFT calculations
also suggest that FAME adsorption through its C¼O group on Mg2þ sites
is much weaker than glycerol adsorption via an OH group on O2 centers.
Therefore, the MgO surface would be mainly covered by glyceroxide
anions that would react with weakly adsorbed FAME molecules.
Downloaded on 16/02/2015 11:50:03.

Acknowledgments
The authors gratefully acknowledge the Universidad Nacional del Litoral
(UNL), Consejo Nacional de Investigaciones Cientı́ficas y Técnicas
(CONICET), and Agencia Nacional de Promoción Cientı́fica y Tecnológica
(ANPCyT), Argentina, for the financial support of this work. They also
thank S. Fuente, R. Ferullo and N. Castellani for their collaboration in
DFT calculations and useful discussions.

References
1 K. Tanabe and K. Saito, J. Catal., 1974, 35, 247.
2 H. Tsuji and H. Hattori, Catal. Today, 2006, 116, 239.
3 A. Corma, S. Iborra, J. Primo and F. Rey, Appl. Catal., 1994, 114, 215.
4 H. Kabashima, H. Tsuji and H. Hattori, Appl. Catal. A: Gen., 1997, 165, 319.
5 S. Bancquart, C. Vanhove, Y. Pouilloux and J. Barrault, Appl. Catal. A : Gen.,
2001, 218, 1.
6 T. F. Dossin, M. F. Reyniers and G. B. Marin, Appl. Catal. B: Environ., 2006,
62, 35.
7 C. A. Ferretti, R. N. Olcese, C. R. Apesteguı́a and J. I. Di Cosimo, Ind. Eng.
Chem Res., 2009, 48, 10387.
8 J. M. Montero, D. R. Brown, P. L. Gai, A. F. Lee and K. Wilson, Chem. Eng. J,
2010, 161, 332.
9 H. Gorzawski and W. F. Hoelderich, J. Mol. Catal. A: Chem., 1999, 144, 181.
10 K. Tanabe, G. Zhang and H. Hattori, Appl. Catal., 1989, 48, 63.
11 H. Kurokawa, T. Kato, T. Kuwabara, W. Ueda, Y. Morikawa, Y. Moro-Oka and
T. Ikawa, J. Catal., 1990, 126, 208.
12 J. I. Di Cosimo, V. K. Dı́ez and C. R. Apesteguı́a, Appl. Catal., 1996, 137, 149.
13 V. K. Dı́ez, C. R. Apesteguı́a and J. I. Di Cosimo, J. Catal., 2006, 240, 235.
14 Y. Xi and R. J. Davis, J. Molec. Catal. A: Chem., 2011, 341, 22.
15 M. Xu, M. J. L. Ginés, A-M. Hilmen, B. L. Stephens and E. Iglesia, J. Catal.,
1997, 171, 130.

26 | Catalysis, 2014, 26, 1–28


View Online

16 J. I. Di Cosimo, C. R. Apesteguı́a, M. J. L. Ginés and E. Iglesia, J. Catal., 2000,


190, 261.
17 T. W. Birky, J. T. Kozlowski and R. J. Davis, J. Catal., 2013, 298, 130.
18 J. J. Ramos, V. K. Dı́ez, C. A. Ferretti, P. A. Torresi, C. R. Apesteguı́a and J. I. Di
Cosimo, Catal. Today, 2011, 172, 41.
19 K. Zhu, J. Hu, C. Kubel and R. Richards, Angew. Chem. Int. Ed., 2006, 45, 7277.
20 H. W Kim, S. H. Shim and J. W. Lee, J. Nanosci. Nanotechnol., 2007, 7, 4434.
21 S. Utamapanya, K. J. Klabunde and J. R. Schlup, Chem. Mater., 1991, 3, 175.
Published on 24 February 2014 on http://pubs.rsc.org | doi:10.1039/9781782620037-00001

22 D. Gulková, O. Solcová and M. Zdrazil, Microporous Mesoporous Mater., 2004,


76, 137.
23 K. T. Ranjit and K. J. Klabunde, Chem. Mater., 2005, 17, 65.
24 D. Chen and E. H. Jordan, Mater. Lett., 2009, 63, 783.
25 H. Hattori, Appl. Catal. A, 2001, 222, 247.
26 B. M. Choudary, K. V. S. Ranganath, U. Pal, M. L. Kantam and B. Sreedhar,
J. Am. Chem. Soc., 2005, 127, 13167.
27 A. O. Menezes, P. S. Silva, E. P. Hernández, L. E. Borges and M. A. Fraga,
Langmuir, 2010, 26, 3382.
28 H. Hattori, Chem. Rev., 1995, 95, 537.
29 P. Kassner and M. Baerns, Appl. Catal. A, 1996, 139, 107.
Downloaded on 16/02/2015 11:50:03.

30 M. León, E. Dı́az, S. Bennici, A. Vega, S. Ordóñez and A. Auroux, Ind. Eng.


Chem. Res., 2010, 49, 3663.
31 G. Busca, Chem. Rev., 2010, 110, 2217.
32 J. L. Fung and I. K. Wang, Appl. Catal. A, 1998, 166, 327.
33 Z. G. Szabó, B. Jóvér and R. Ochmacht, J. Catal., 1975, 39, 225.
34 H. Kurokawa, T. Kato, T. Kuwabara, W. Ueda, Y. Morikawa, Y. Moro-Oka and
T. Ikawa, J. Catal., 1990, 126, 208.
35 V. K. Dı́ez, C. R. Apesteguı́a and J. I. Di Cosimo, J. Catal., 2003, 215, 220.
36 A. Auroux, P. Artizzu, I. Ferino, R. Monaci, E. Rombi, V. Solinas and
G. Petrini, J. Chem. Soc., Faraday Trans., 1996, 92, 2619.
37 S. Kús, M. Otremba, A. Tórz and M. Taniewski, Appl. Catal. A: Gen., 2002, 230, 263.
38 H. Lauron-Pernot, F. Luck and J. M. Popa, Appl. Catal., 1991, 78, 213.
39 M. Huang and S. Kaliaguine, Catal. Lett., 1993, 23, 373.
40 P. Thomasson, O. S. Tyagi and H. Knözinger, Appl. Catal. A: Gen., 1999, 181, 181.
41 A. Gervasini, J. Fenyvesin and A. Auroux, Catal. Lett., 1997, 43, 219.
42 S. V. Bordawekar, E. J. Doskocil and R. Davis, Catal. Lett., 1997, 44, 193.
43 V. K. Dı́ez, C. R. Apesteguı́a and J. I. Di Cosimo, Catal. Today, 2000, 63, 53.
44 M. A. Aramendı́a, V. Borau and C. Jiménez, A. Marinas, J. M. Marinas, J. R.
Ruiz and F. J. Urbano, J. Molec. Catal. A: Chem., 2004, 218, 81.
45 C. Arrouvel, M. Digne, M. Breysse, H. Toulhoat and P. Raybaud, J. Catal,.,
2004, 222, 152.
46 C. Chizallet, G. Costentin, M. Che, F. Delbecq and P. Sautet, J. Phys. Chem. B,
2006, 110, 15878.
47 H. Petitjean, K. Tarasov, F. Delbecq, P. Sautet, J. M. Krafft, P. Bazin, M. C.
Paganini, E. Giamello, M. Che, H. Lauron-Pernot and G. Costentin, J. Phys.
Chem. C, 2010, 114, 3008.
48 D. Cornu, H. Guesmi, J. M. Krafft and H. Lauron-Pernot, J. Phys. Chem. C,
2012, 116, 6645.
49 C. A. Ferretti, A. Soldano, C. R. Apesteguı́a and J. I. Di Cosimo, Chem. Eng. J.,
2010, 161, 346.
50 M. L. Bailly, C. Chizallet, G. Costentin, J. M. Krafft, H. Lauron-Pernot and
M. Che, J. Catal., 2005, 235, 413.
51 R. Vidruk, M. V. Landaua, M. Herskowitz, M. Talianker, N. Frage, V. Ezersky
and N. Froumin, J. Catal., 2009, 263, 196.

Catalysis, 2014, 26, 1–28 | 27


View Online

52 V. K Dı́ez, C. A. Ferretti, P. A. Torresi, C. R. Apesteguı́a and J. I. Di Cosimo,


Catal. Today, 2011, 173, 21.
53 J. I. Di Cosimo, V. K. Dı́ez, M. Xu, E. Iglesia and C. R. Apesteguı́a, J. Catal.,
1998, 178, 499.
54 C. Morterra, G. Ghiotti, F. Boccuzzi and S. Coluccia, J. Catal., 1978, 51, 299.
55 R. Philipp and K. Fujimoto, J. Phys. Chem., 1992, 96, 9035.
56 F. Prinetto, G. Ghiotti, R. Durand and D. Tichit, J. Phys. Chem. B, 2000, 104, 11117.
57 M. León, E. Dı́az, S. Bennici, A. Vega, S. Ordóñez and A. Auroux, Ind. Eng.
Published on 24 February 2014 on http://pubs.rsc.org | doi:10.1039/9781782620037-00001

Chem. Res., 2010, 49, 3663.


58 E. Knozinger, K. Jacob, S. Singh and P. Hofmann, Surf. Sci, 1993, 290, 388.
59 M. M. Branda, R. M. Ferullo, P. G. Belelli and N. J. Castellani, Surf. Sci., 2003,
527, 89.
60 H. Tsuji, T. Shishido, A. Okamura, Y. Gao, H. Hattori and H. Kita, J. Chem.
Soc. Faraday Trans., 1994, 90, 803.
61 J. V. Evans and T. L. Whateley, Trans. Faraday Soc., 1967, 63, 2769.
62 V. K. Dı́ez, J. I. Di Cosimo and C. R. Apesteguı́a, Appl. Catal. A: Gen., 2008,
345, 143.
63 E. Brenna, C. Fuganti, S. Serra and P. Kraft, Eur. J. Org. Chem., 2002, 967.
64 Rhodia Inc., P. Gradeff, US Patent, 3 840 601, 1974
Downloaded on 16/02/2015 11:50:03.

65 Union Camp Corporation, P. Mitchell, US Patent, 4 874 900, 1989


66 J. C. Roelofs, A. J. van Dillen and K. P. de Jong, Catal. Today, 2000, 60, 297.
67 S. Abelló, S. Dhir, G. Colet and J. Pérez-Ramı́rez, Appl. Catal. A: Gen., 2007,
325, 121.
68 C. Noda Pérez, C. A. Henriques, O. A. C. Antunes and J. L. F. Monteiro, J. Mol.
Catal. A: Chem., 2005, 223, 83.
69 V. Ponec, Appl. Catal. A, 1997, 149, 27.
70 P. Claus, Appl. Catal. A: Gen., 2005, 291, 222.
71 R. L. Augustine, Catal. Today, 1997, 37, 419.
72 C. F. de Graauw, J. A. Peters, H. van Bekkum and J. Huskens, Synthesis, 1994,
1007.
73 M. G. Musolino, P. De Maio, A. Donato and R. Pierpaolo, J. Mol. Catal. A:
Chem., 2004, 208, 219.
74 G. Szollosi and M. Bartok, J. Mol. Catal. A: Chem., 1999, 148, 265.
75 A. Corma, M. E. Domine and S. Valencia, J. Catal., 2003, 215, 294.
76 J. I. Di Cosimo, A. Acosta and C. R. Apesteguı́a, J. Mol. Catal. A: Chem., 2004,
222, 87.
77 J. I. Di Cosimo, A. Acosta and C. R. Apesteguı́a, J. Mol. Catal. A: Chem., 2005,
234, 111.
78 Y. Zheng, X. Chen and Y. Shen, Chem. Rev., 2008, 108, 5253.
79 A. Corma, S. B. A. Hamid, S. Iborra and A. Velty, J. Catal., 2005, 234, 340.
80 N. O. V. Sonntag, J. Amer. Oil Chem. Soc., 1982, 59, 795A.
81 J. Barrault, S. Bancquart and Y. Pouilloux, C. R. Chim., 2004, 767, 593.
82 C. A. Ferretti, C. R. Apesteguı́a and J. I. Di Cosimo, App. Catal. A: Gen., 2011,
399, 146.
83 C. A. Ferretti, S. Fuente, R. Ferullo, N. Castellani, C. R. Apesteguı́a and J. I. Di
Cosimo, Appl. Catal. A: Gen., 2012, 413–414, 322.
84 A. D. Becke, J. Chem. Phys., 1993, 98, 5648.
85 M. Calatayud, A. M. Ruppert and B. M. Weckhuysen, Chem. Eur. J, 2009, 15,
10864.
86 M. M. Branda, A. H. Rodrı́guez, P. G. Belelli and N. J. Castellani, Surf. Sc.,
2009, 603, 1093.
87 A. E. Reed, L. A. Curtiss and F. Weinhold, Chem. Rev., 1988, 88, 899.
88 T. Trevethan and A. L. Shluger, J. Phys. Chem. C, 2007, 111, 15375.

28 | Catalysis, 2014, 26, 1–28

You might also like