• Non ci sono risultati.

The Influence of Alcohols in Driving the Morphology of Magnesium Chloride Nanocrystals

N/A
N/A
Protected

Academic year: 2021

Condividi "The Influence of Alcohols in Driving the Morphology of Magnesium Chloride Nanocrystals"

Copied!
8
0
0

Testo completo

(1)

This is an author version of the contribution published on:

Questa è la versione dell’autore dell’opera:

[Thushara et al., 9, ChemCatChem, 2017, pagg 1782–1787]

The definitive version is available at:

La versione definitiva è disponibile alla URL:

(2)

The Influence of Alcohols in Driving the Morphology of MgCl

2

Nanocrystals

K. S. Thushara, Maddalena D’Amore*, Alessandro Piovano, Silvia Bordiga and Elena Groppo*

Abstract: MgCl2 nano-crystals prepared in the presence of methanol

and ethanol have been characterized by complementing surface science tools with a systematic quantum-mechanical investigation of the stability order (in terms of Gibbs free energy) of different MgCl2

surfaces in the presence of the two alcohols. Both the alcohols drastically change the overall stability and the stability order of the exposed surfaces, mainly as a consequence of the entropic contribution to the Gibbs free energy, hence inhibiting or promoting crystal growth in certain directions. The environment-dependent

occurrence of the MgCl2 surfaces may influence the structure and

the properties of the supported TiClx sites in the MgCl2/TiCl4

pre-catalyst, with important implications in the design of morphologically controlled Ziegler-Natta catalysts.

Many efficient and selective catalytic materials for a wide range of energy and environmental applications are composed of nanometer-sized particles exposing different surface sites.[1-5]

Identifying the correlation between the particle structure at the nano-scale and their activity and selectivity is fundamental to develop tailored catalysts, a challenge that up to now has been almost a privilege of homogeneous catalysts. However, the experimental characterization of the surface and structural properties of nano-catalysts is problematic,[6-10] especially in

reaction conditions, and in most of the cases the obtained information is averaged over the whole ensemble of sites. Recently, a huge effort has being devoted to synthetize nano-structures with controlled morphology, where well-defined

exposed surfaces display a specific catalytic behaviour.[11-13] The

use of surfactants to stabilize certain surfaces with respect to others is a well-established and efficient method to drive the morphology of nanomaterials in general, ranging from metal-oxides to metal nanoparticles.[11,14-19] The design of

nano-catalysts with a defined morphology requires the optimization of the synthetic protocol, i.e. of the thermodynamic and environmental conditions of operation. In this process, the theoretical development of accurate adsorption models may help. In this respect, the computational design of heterogeneous catalysts is becoming feasible, thanks to the recent advances in Density Functional Theory (DFT), which makes it possible to

accurately describe the surface processes.[20]

In this work we show the potential of combining an accurate surface characterization with state-of-the-art DFT methods for

the prediction of the structural and surface properties of MgCl2

nano-crystals prepared in the presence of different alcohols,

which are relevant components in Ziegler-Natta (ZN) catalysis. Modern heterogeneous ZN catalysts for olefin polymerization may be seen as stereotypes of complex nanometre-sized materials. Regarding the complexity, it mainly derives from their multi-component nature,[21,22] consisting of MgCl

2, TiCl4, an

aluminium alkyl (e.g. AlEt3), and organic Lewis Bases (LB, e.g.

aromatic mono- and di-esters), the latter introduced to achieve a high stereo-selectivity in propene polymerization. These systems can be definitely labelled – and treated with all the related techniques – as nanometer-sized materials, since the MgCl2

activated support is featured with a relevant structural disorder along the c axis and nanometric crystalline dimensions.[23-26]

Disordered MgCl2 is usually synthesized in the presence of TiCl4

following either mechanical or chemical routes.[27-30] In

mechanical methods, MgCl2 and TiCl4 are co-milled with or

without the solvent to generate the active catalyst. In chemical

routes, highly crystalline MgCl2 (α form) is interacted with Lewis

bases (mainly alcohols), then de-alcoholated or reacted with TiCl4 to form disordered and nano-sized MgCl2 ( form). The

influence of the synthesis protocol in determining the structural, morphological and surface properties of the final catalyst is well known in the industrial practice.[31] It is recognized that the nature

of the molecular adduct and the alcohol employed in the synthesis of MgCl2 drastically influence the porosity of the

catalyst and hence its activity and the properties of the obtained polymer.[31-33] “Super active” catalyst supports have been

prepared from molecular adducts between MgCl2 and alcohols.

[34,35] However, a clear correlation between the synthesis

variables and the properties of the catalyst at the nano-scale is still missing.

Since the 1980s,[36,37] ZN catalysts have been traditionally

investigated by means of molecular mechanics and quantum mechanical computations, aimed at identifying possible reaction mechanisms for olefin polymerization on what were supposed to

be the most stable MgCl2 surfaces in the presence of LB and

TiClx. In the last decade, the application of DFT and DFT

methods including dispersion (DFT-D) to elucidate the energetic of interaction and the local structure of TiCl4 (or TixCl3x) and LB

at these surfaces[25,38-46] substantially renewed the seminal vision

of catalysis on MgCl2. Recent simulations performed on clusters

of different size and morphology[41,42] demonstrated that the

relative stability of MgCl2 surfaces drastically changes in the

presence of donors. This holds for several small donor molecules (H2O,[44] NH3,[44] methanol,[42,44] ethanol[44]), as well as

for the industrially employed donors.[44,46-50] Interestingly, these

theoretical studies validated the experimental results obtained years before by Thune and co-workers on a planar model of MgCl2,[51,52] when it was visualized for the first time the effect of

an internal donor in controlling the morphology of the MgCl2

crystals formed during the preparation of a ZN catalyst. Although so far all the computational studies were performed only on the (110) and (104) surfaces (the most plausible ones in naked

MgCl2 crystals), these results clearly indicate that the reaction

environment affects the free energy of the exposed MgCl2

Dr. K. S. Thushara, Dr. M. D’Amore, A. Piovano, Prof. S. Bordiga, Dr. E. Groppo

Department of Chemistry, INSTM and NIS Centre University of Torino

Via Quarello 15, 10135 Torino, Italy

E-mail: elena.groppo@unito.it, maddalena.damore@unito.it

Supporting information for this article is given via a link at the end of the document.

(3)

surfaces. Hence, the MgCl2 nano-crystals are expected to have

a dynamic behaviour (i.e. to change the overall morphology) during the catalyst synthesis, and likely also during the catalyst function.

Recently, we accomplished the first systematic

computational investigation of the most stable surfaces of MgCl2

aiming at obtaining their stability order in terms of Gibbs free energy both for naked crystals and in the presence of methanol.

[53] Our computational morphological analysis revealed that the

surfaces exposing highly acidic Mg2+ cations (namely the (012),

(015), and (110) surfaces) are highly stabilized by methanol, especially at the temperature typically adopted in the preparation of the pre-catalysts. We also demonstrated that FT-IR spectroscopy of CO adsorbed at 100 K, coupled with vibrational analysis performed on adducts CO/surface (hkl), is a powerful method to characterize the surface of structurally disordered MgCl2.[53] In the present work we complete our previous

systematic study by investigating the effect of ethanol in driving the morphology of MgCl2 nano-crystals, in the attempt to

rationalize the role of alcohols as shape-directing agents for

MgCl2 crystals. We wish to underline that our work constitutes a

step further most of the previous theoretical works describing the

adsorption of donors on the MgCl2 surfaces, that do concentrate

only on the “classical” (104) and (110) surfaces (whose presence was inferred on the basis of the pure crystallographic perfection) and give an estimation of the adsorption enthalpy only, neglecting the entropic term.

Briefly, the disordered MgCl2 nano-crystals were obtained

upon controlled de-alcoholation of the corresponding MgCl2·6ROH adducts,[53-55] where R = CH3 (Me) or C2H5 (Et),

resulting in MgCl2-MeOH and MgCl2-EtOH samples, respectively.

According to the thermogravimetric data shown in Figure 1a, all the alcohol molecules are removed below 200 °C, although the de-alcoholation process proceeds with different steps in the two

cases. Particularly, the MgCl2·6EtOH adduct loses only four out

of the six ligands below 150 °C, and the last two in the 150-200 °C range, in very good agreement with previously

Figure 1. Part a): Thermogravimetric (full lines) and differential

thermogravimetric (dotted lines) analysis of the two alcohol precursors, MgCl 2-6EtOH (black) and MgCl2-6MeOH (grey). Part b): XRPD patterns of the resulting disordered MgCl2 nano-crystals: MgCl2-EtOH (black) and MgCl2-MeOH (grey). The two patterns are vertically translated for clarity. The broad peak centered around 2θ = 20° is due to the glass capillary.

published data.[56] In contrast, at 150 °C the MgCl

2·6MeOH

adduct has already lost five out of the six ligands, in agreement with the higher volatility of methanol with respect to ethanol. In

both cases, a layered MgCl2 phase characterized by an

extensive structural disorder is obtained, as demonstrated by the X-Ray powder diffraction (XRPD) patterns shown in Figure 1b and in agreement with literature.[55,57-60] The two diffraction

patterns are very similar, irrespective of the starting precursor. This is a clear limit of the XRPD technique, which is unable to reveal subtle structural differences for nano-sized and disordered materials.

The surface properties of the two MgCl2 samples were

determined by combining N2 physisorption measurements with

FT-IR spectroscopy of CO adsorbed at 100 K. The former technique gives indication on the total specific surface area, while the latter discriminate the type of exposed surface sites. Indeed, we have recently demonstrated that this technique can

be successfully used to probe the exposed MgCl2 surfaces in

terms of coordinatively unsaturations and polarizing ability of the

exposed Mg2+ cations, and hence of their Lewis acidity.[53] Both

samples show adsorption isotherms of Type IV (Figure 2a) and are characterized by a pronounced hysteresis loop, which indicates the presence of inter-particles mesopores. The specific surface area, evaluated by means of the BET model, is of about

40 m2/g and 100 m2/g for MgCl

2-EtOH and MgCl2-MeOH, respectively.

The FT-IR spectra, in the ν̃(CO) region, of CO adsorbed at

100 K on the MgCl2-EtOH and MgCl2-MeOH as a function of the CO

coverage (θCO) are reported in Figure 2b and Figure 2c,

respectively. Two absorption bands are observed in both cases, having the same behavior as a function of θCO, but with a

different total and relative intensity. These two bands have been assigned in our previous work based on a detailed vibrational analysis.[53] Briefly, the absorption band located at 2182 cm-1 for

maximum CO and shifting to 2194 cm-1 for CO → 0, was

interpreted as the superposition of three contributions, namely CO adsorbed on the (110), (015) and (012) surfaces. On these

surfaces the Mg2+ cations are either tetra-coordinated (on the

(110) face) or penta-coordinated but with a high polarizing ability due to the proximity of chlorine ligands belonging to a

neighbouring layer. Hereafter we will refer to these Mg2+ sites

with the name of “strongly acidic Mg2+ sites”. The second

absorption band, located at 2163 cm-1 at maximum  CO and

much less sensitive to the decrease of coverage, was assigned to CO adsorbed on the (104) and (107) penta-coordinated

families of surfaces. We will call these sites “weakly acidic Mg2+

sites”. No absorption bands due to CO adsorbed on defects

(expected around 2200 cm-1) are observed, likely a

consequence of the persistence of a minor amount of alcohol molecules (or some by-products)[61,62] on the most energetic

surface sites. The absolute intensity of the ν̃(CO) absorption

bands at the maximum θCO is smaller for MgCl2-EtOH (Figure 2b)

than for MgCl2-MeOH (Figure 2c), in fair agreement with the

specific surface area determined by the BET analysis (Figure 2a). More important, the relative intensity of the two absorption

bands is reversed in the two cases. For CO adsorbed on MgCl

2-EtOH (Figure 2b), the absorption band assigned to weakly acidic

(4)

Figure 2. Part a): N2 adsorption isotherms at 77 K for MgCl2-EtOH (black) and MgCl2-MeOH (grey). Parts b) and c): FT-IR spectra (in the ν̃(CO) region) of CO adsorbed at 100 K on MgCl2-EtOH and MgCl2-MeOH, respectively, as a function of the CO coverage (θmax in bold). Part d): as part c) but for a MgCl2-MeOH sample de-alcoholated at 300 °C. All the FT-IR spectra are normalized to the thickness of the pellet and subtracted to the spectrum collected before CO dosing.

strongly acidic Mg2+ cations, while the opposite behavior is

observed for CO adsorbed on MgCl2-MeOH (Figure 2c). Hence,

despite MgCl2-EtOH and MgCl2-MeOH are almost indistinguishable by

XRPD, they clearly expose different surfaces, at a different extent. This is the first experimental proof of the role of alcohol in

driving the morphology of MgCl2 at the nano-scale, so far only

predicted by theoretical calculations.[43] It is also interesting to

notice that the relative intensity of the two ν̃(CO) absorption bands does not change when the MgCl2 nano-crystals are

obtained at higher de-alcoholation temperature. For example, Figure 2d shows the FT-IR spectra of CO adsorbed at 100 K on a MgCl2-MeOH samples obtained upon de-alcoholation of the

corresponding adduct at 300 °C. With respect to MgCl2-MeOH

de-alcoholated at 200 °C, only the absolute intensity of the spectra decreases, while the relative intensity of the two ν̃(CO) absorption bands remains almost the same. Similar results were obtained for MgCl2-EtOH. This clearly indicates that the

morphology of the MgCl2 nano-crystals is strictly dependent on

the type of alcohol. In contrast, the activation temperature seems to influence the total surface area (i.e. the dimension of the crystals), in that a high activation temperature likely induces sintering phenomena.

In the attempt to rationalize the behaviour of MgCl2 in the

presence of different alcohols we extended our systematic computational study on the stability of the exposed surfaces, previously performed in the presence of methanol, to the ethanol case. The results are summarized in Table 1. A graphical representation of the computed surfaces in the presence of methanol can be found in our previous work.[53] Those with

ethanol, except for longer alkyl chains, are practically indistinguishable,. Briefly, we considered the most stable surfaces of α-MgCl2 and for all of them we evaluated the

variation of the total electron energy due to the adsorption of

alcohol molecules (ΔEads), as well as the Gibbs free energy of

adsorption (ΔGads) at 298 K. The calculations were performed

considering an alcohol coverage ROH = 1 for all the surfaces,

except for the (110) one exposing tetra-coordinated Mg2+

cations, where also ROH = 2 was considered. Table 1 reports

also the Gibbs free energy (which is a measure of the stability)

of each surface in the presence of alcohol in the gas (Gs-ROH(g))

and liquid (Gs-ROH(l)) phase (in J/m2), calculated by adding ΔGads to

the formation energy of each naked surface starting from bulk crystal (Gs), i.e. Gs-ROH = Gs + ΔGads. Finally, the calculations were

performed also at 373 K. Indeed, the de-alcoholation of the

MgCl26ROH adduct and the formation of the activated MgCl2

nano-crystals occurs in the 298 – 473 K temperature range. Also

the typical processes for the preparation of MgCl2 with improved

controlled morphology starting from MgCl2 alcoholate adducts

are conducted around 373 K.

According to DFT-D calculations, in the presence of adsorbates a huge stabilization of all the surfaces occurs. In particular, both alcohols selectively stabilize the (110) surface with respect to naked MgCl2, in agreement with the recent

literature.[40-44] Our systematic computational approach allows

appreciating relevant differences between methanol and ethanol.

In particular, the Gibbs free energy of adsorption (ΔGads in Table

1) of ethanol on the (012), (101) and (110) surfaces are substantially lower than those of methanol on the same surfaces, which means that the (012), (101) and (110) families of surfaces are less stabilized by ethanol with respect to methanol. More in details, for the (012) and (101) surfaces, the enthalpic contribution already disfavours the ethanol adsorption with

respect to methanol (lower ΔEads), while for the (110) surface

entropic reasons hinder the adsorption of ethanol (similar ΔEads

for methanol and ethanol). At the same time, ΔGads of ethanol on

the (104) and (107) families of surfaces are higher than those of methanol, indicating that these surfaces are stabilized more by ethanol than by methanol. In this case the phenomenon is driven mainly by the entropic contribution. Finally, the stability of the (015) surface is the same, irrespective of the alcohol. The

stability (Gs) of the different surfaces is a balance between the

enthalpic and the entropic contributions, and follows a different order in the presence of the two alcohols.

Table 1. Predicted adsorption energies (ΔEads and ΔGads) for ethanol and methanol on different MgCl2 surfaces, and predicted stability (in terms of Gibbs free energy, in J/m2) of each surface in the presence of ethanol and methanol in the gas phase (Gs-ROH(g)) and in the liquid phase (Gs-ROH(l)) at 298 K and 373 K.

Surface (hkl)

R-OH ΔEads

(kJ/mol, 298K) ΔGads (kJ/mol, 298K) Gs-ROH(g) (J/m2, 298 K) Gs-ROH(l) [a] (J/m2, 298 K) Gs-ROH(g) (J/m2,373.15 K)

(5)

EtOH MeOH EtOH MeOH EtOH MeOH EtOH MeOH EtOH MeOH

(001) 1 -34.9 - 35.4 - 0.0498 0.0498 0.0429 (101) 1 -113.5 -122.1 -48.0 -51.3 0.236 0.210 0.283 0.243 0.408 0.381 (012) 1 -116.6 -136.6 -50.3 -63.1 0.188 0.0947 0.233 0.126 0.340 0.249 (104) 1 -83.8 -83.8 -20.0 -15.2 0.556 0.586 0.595 0.613 0.730 0.773 (015) 1 -101.8 -100.9 -34.4 -34.8 0.259 0.257 0.294 0.282 0.388 0.370 (107) 1 -84.2 -87.1 -24.2 -18.9 0.271 0.297 0.300 0.317 0.383 0.410 (110) 1 -116.7 -118.2 -60.6 -52.8 0.637 0.672 0.665 0.717 0.813 0.850 (110) 2 -109.5 -111.5 -27.4 -42.8 0.513 0.479 0.586 0.518 0.655 0.713

[a] Thermodynamic data of normal alcohols as reported in references [63-65].

The computational results summarized in Table 1 allow explaining the different relative intensity of the two ν̃(CO) absorption bands observed in the FT-IR spectra of CO adsorbed at 100 K on MgCl2-EtOH and MgCl2-MeOH. As a matter of fact,

calculations predict that ethanol, more than methanol, stabilizes

the surfaces exposing weakly acidic Mg2+ cations (i.e. the (104)

and (107) surfaces), which are responsible of the absorption band at 2163 cm-1. At the same time, most of the surfaces

exposing strongly acidic Mg2+ sites (i.e. the (110) and (012)),

which account for the absorption band at 2182 cm-1, are less

stabilized by ethanol than by methanol. On the contrary, this computational approach does not allow predicting the total intensity of the FT-IR spectra in the ν̃(CO) region, that means the total specific surface area of MgCl2 nano-crystals. In this

respect it is important to notice that our approach allows determining which are the possible surfaces in the chosen experimental conditions (donors, temperature). However, the degree of expression of the possible surfaces depends on the crystallization kinetic, which clearly influences also the

dimension of the MgCl2 nano-crystals.

The experimental and computational results reported above demonstrate that a careful vibrational analysis conducted at DFT-D level, coupled with FT-IR spectroscopy of CO adsorbed at 100 K, is a very powerful tool for predicting and validating the morphology of disordered MgCl2 nano-crystals formed in the

presence of different alcohols as Lewis bases. We have demonstrated that the alcohol drastically changes the overall stability and the stability order of the exposed surfaces, mainly as a consequence of the entropic contribution to the Gibbs free energy. Our results definitely prove that long chain aliphatic alcohols would stabilize those surfaces exposing weakly acidic Mg2+ cations, while methanol favors the surfaces exposing

strongly acidic Mg2+ sites, irrespective of their crystallographic

definition in terms of (hkl) indexes, which is sometimes criticized when dealing with nano-sized materials. Since the distribution

and the extension of the MgCl2 surfaces influence the structure

and the properties of the supported TiClx sites in the MgCl2/TiCl4

pre-catalyst, these results are of primary interest in the design of morphologically controlled ZN catalysts.

Experimental and Computational Details Section

The MgCl2·6ROH adducts were synthesized as reported in refs. [53-55].

The active MgCl2-ROH phases was obtained from the corresponding

adducts through a controlled de-alcoholation in dynamic vacuum at 473 K for a prolonged time, as described in ref. [53]. The thermal gravimetric

analysis (TGA) was carried out on a TAQ600 instrument, at a heating ramp rate of 1 °C/min from room temperature to 300 °C in nitrogen flow, after an equilibration in N2 at room temperature at 30 °C for 30 minutes.

X-ray Diffraction (XRPD) patterns were collected in the capillary configuration with a PW3050/60 X’Pert PRO MPD diffractometer from PANalytical working in Debye−Scherrer geometry, using as a source the high power ceramic tube PW3373/10 LFF with a Cu anode, equipped with a Ni filter to attenuate Kβ and focused by X-ray mirror PW3152/63. Scattered photons were collected by a RTMS (real time multiple strip) X’celerator detector. The surface area of de-alcoholated MgCl2 samples

was determined by the Brunauer–Emmett–Teller (BET) method via nitrogen adsorption ay 77 K performed on a ASAP 2020 instrument. In all the cases, care was taken to avoid exposure of the samples to air. A Bruker Vertex70 spectrophotometer equipped with a MCT detector was used to collect in-situ FT-IR spectra of disordered MgCl2-ROH in the

presence of CO at 100 K, as a function of CO coverage. The standard procedure is described in reference [53].

Density Functional calculations were performed with the CRYSTAL14 code employing a Gaussian type basis set.[66] The hybrid Becke, three

parameters, Lee-Yang-Parr (B3LYP) functional was adopted with a parametrized damped 1/r6 term added to the energy obtained with the

standard density functional methods to account for dispersion. The DFT-D2 version proposed by Grimme[67] with a new parametrization for

crystals was adopted.[68] The inclusion of dispersion terms proper of

DFT-D formulations allows to overcome the intrinsic limit of standard DFT-DFT functionals lacking long-range contributions. All - electron Gaussian-type Split valence triple- basis sets plus polarization (TZVP) functions were applied to describe all the elements (Mg, Ti and Cl atoms).[69] An Ahlrichs

VTZ[70] plus polarization quality was adopted for the adsorbed alcohol

molecules, whereas the coefficients of the polarization Gaussian functions (αpol) were optimized in preliminary studies, thus enabling more

accurate calculations of the adsorption energy and vibrational frequencies. The choice was aimed at reducing the Basis Set Superposition Error (BSSE), that may become quite large in particular when the dispersion correction is included during the optimization process, i.e. when the molecule and the surface become closer. In a previous paper by us on methanol adsorption at the same level of computations we computed the BSSEs for all the surfaces and their values fell in the range 9.4%÷12% of the computed total adsorption energy, indicating that the ‘customized’ Ahlrichs basis sets adopted are adequate. Due to the low values of the correction in the adsorption energies (both ΔEads and ΔGads), in this work we did not included the

BSSE corrections. The Gauss-Legendre quadrature and Lebedev schemes were used to generate angular and radial points of a pruned grid consisting of 75 radial points and maximum number of 974 angular points over which electron density and its gradient were integrated. Values of the tolerance that control the Coulomb and exchange series in periodical systems were set to 7 7 7 7 18. All the bielectronic integrals, Coulomb and exchange, were evaluated exactly. For all B3LYP-D calculations, 10 K points have been adopted. For all the surfaces, calculations were performed considering a methanol coverage of θMeOH =

1, and for the (110) and (015) surfaces also θMeOH = 2 was considered,

although a second methanol molecule at this level of computation seems to be stable only on the (110) surface. By means of the “Planes” tool implemented in CRYSTAL code, the analysis of the -MgCl2 planes was

performed to locate the most stable surfaces of the crystals following the Bravais' law, since the surface free energy is inversely proportional to the directional density and thus proportional to dhkl (i.e the distance between

(6)

the planes).The {hkl} planes that have the interplanar distance in the range (� �hkl, �hklmax) have been analysed, (where �hklmax is the maximum

interplanar distance for a given plane {hkl}) and f is a factor between 0 and 1, in our case all plane families were considered up to f=0.3.

Acknowledgements

Adriano Zecchina and Gabriele Ricchiardi are kindly acknowledged for the useful discussion. The computational work was performed on both the Abel Cluster, owned by the University of Oslo and the Norwegian meta-centre for High Performance Computing (NOTUR), and operated by the Department for Research Computing at USIT, the University of Oslo IT-department (http://www.hpc.uio.no/) and at the HPC centre CINECA, thanks to the Galileo resource accessed through the POLCAT ISCRA C Project. This work has been supported by the Progetto di Ateneo/CSP 2014 (Torino_call2014_L1_73).

Keywords: MgCl2• nanocrystals • DFT-D calculations • FT-IR spectroscopy• Ziegler-Natta catalysts

[1] Catal. Lett. 2001, 76, 111-124.

[2] R. Schlogl, S. B. Abd Hamid, Angew. Chem.-Int. Edit. 2004, 43, 1628-1637. [3] D. Astruc, F. Lu, J. R. Aranzaes, Angew. Chem.-Int. Edit. 2005, 44,

7852-7872.

[4] A. Zecchina, E. Groppo, S. Bordiga, Chem. Eur. J. 2007, 13, 2440 – 2460. [5] R. Ye, T. J. Hurlburt, K. Sabyrov, S. Alayoglu, G. A. Somorjai, Proc. Natl.

Acad. Sci. U. S. A. 2016, 113, 5159-5166.

[6] E. Gross, G. A. Somorjai, J. Catal. 2015, 328, 91-101.

[7] S. Alayoglu, J. M. Krier, W. D. Michalak, Z. Zhu, E. Gross, G. A. Somorjai,

ACS Catal. 2012, 2, 2250-2258.

[8] I. L. C. Buurmans, B. M. Weckhuysen, Nat. Chem. 2012, 4, 873-886. [9] L. C. Gontard, L.-Y. Chang, C. J. D. Hetherington, A. I. Kirkland, D.

Ozkaya, R. E. Dunin-Borkowski, Angew. Chem. Int. Ed. 2007, 46, 3683-3685.

[10] H. Topsøe, J. Catal. 2003, 216, 155-164.

[11] I. Lee, R. Morales, M. A. Albiter, F. Zaera, Proc. Natl. Acad. Sci. U. S. A.

2008, 105, 15241-15246.

[12] I. Lee, F. Delbecq, R. Morales, M. A. Albiter, F. Zaera, Nat. Mater. 2009,

8, 132-138.

[13] M. G. Taylor, N. Austin, C. E. Gounaris, G. Mpourmpakis, ACS Catal.

2015, 5, 6296-6301.

[14] P. J. F. Harris, Nature 1986, 323, 792-794.

[15] Z. R. Tian, J. A. Voigt, J. Liu, B. McKenzie, M. J. McDermott, J. Am.

Chem. Soc. 2002, 124, 12954-12955.

[16] Z. R. Tian, J. A. Voigt, J. Liu, B. McKenzie, M. J. McDermott, M. A. Rodriguez, H. Konishi, H. Xu, Nat. Mater. 2003, 2, 821-826.

[17] Y. Yin, A. P. Alivisatos, Nature 2005, 437, 664-670.

[18] L. P. Hansen, Q. M. Ramasse, C. Kisielowski, M. Brorson, E. Johnson, H. Topsøe, S. Helveg, Angew. Chem. Int. Ed. 2011, 50, 10153-10156. [19] J. Anton, J. Nebel, H. Song, C. Froese, P. Weide, H. Ruland, M. Muhler,

S. Kaluza, J. Catal. 2016, 335, 175-186.

[20] J. K. Norskov, T. Bligaard, J. Rossmeisl, C. H. Christensen, Nat. Chem.

2009, 1, 37-46.

[21] L. Cavallo, S. Del Piero, J.-M. Ducéré, R. Fedele, A. Melchior, G. Morini, F. Piemontesi, M. Tolazzi, J. Phys. Chem. C 2007, 111, 4412-4419. [22] E. Groppo, K. Seenivasan, C. Barzan, Catal. Sci. Technol. 2013, 3,

858-878.

[23] U. Giannini, Makromol. Chem. Suppl. 1981, 5, 216.

[24] R. Zannetti, C. Marega, A. Marigo, A. Martorana, J. Polym. Sci. B 1988,

26, 2399-2412.

[25] V. Busico, M. Causa, R. Cipullo, R. Credendino, F. Cutillo, N. Friederichs, R. Lamanna, A. Segre, V. V. Castellit, J. Phys. Chem. C 2008, 112, 1081-1089.

[26] T. Taniike, M. Terano, J. Catal. 2012, 293, 39-50. [27] F. J. Karol, Catal. Rev. Sci. Eng. 1984, 26, 557.

[28] F. J. Karol, K. J. Cann, B. E. Wagner, in: Transition Metals and Organometallics as Catalysts for Olefin Polymerization; Kaminsky, W.; Sinn, H., Ed.; Springer: Berlin, Germany, 1988, p. 149.

[29] Y. V. Kissin, in: Alkene Polymerization Reactions with Transition Metal Catalysts; Elsevier: Amsterdam, The Netherlands, 2008.

[30] T. E. Nowlin, R. I. Mink, Y. V. Kissin, in: Transition Metal Polymerization Catalysts; Hoff, R.; Mathers, R. T., Ed.; John Wiley & Sons: Hoboken, USA., 2009.

[31] E. Albizzati, U. Giannini, G. Collina, L. Noristi, L. Resconi, Catalysts and

polymerizations, in: Polypropylene Handbook; Moore, E. P. J., Ed.;

Hanser-Gardner Publications: Cincinnati, OH, 1996.

[32] T. Taniike, P. Chammingkwan, V. Q. Thang, T. Funako, M. Terano, Appl.

Catal. A.: General 2012, 437–438, 24-27.

[33] T. Taniike, T. Funako, M. Terano, J. Catal. 2014, 311, 33-40.

[34] K. S. Thushara, R. Mathew, T. G. Ajithkumar, P. R. Rajamohanan, S. Bhaduri, C. S. Gopinath, J. Phys. Chem. C 2009, 113, 8556-8559. [35] J. C. W. Chien, Y. Hu, J. C. Vizzini, J. Polym. Sci. A 1990, 28, 273-284. [36] P. Corradini, V. Barone, R. Fusco, G. Guerra, Gazz. Chim. Ital. 1983, 113,

601.

[37] P. Corradini, G. Guerra, L. Cavallo, Acc. Chem. Res. 2004, 37, 231-241. [38] M. Boero, M. Parrinello, K. Terakura, J. Am. Chem. Soc. 1998, 120,

2746-2752.

[39] M. Boero, M. Parrinello, H. Weiss, S. Huffer, J. Phys. Chem. A 2001, 105, 5096-5105.

[40] R. Credendino, V. Busico, M. Causa, V. Barone, P. H. M. Budzelaar, C. Zicovich-Wilson, Phys. Chem. Chem. Phys. 2009, 11, 6525-6532. [41] R. Credendino, J. T. M. Pater, A. Correa, G. Morini, L. Cavallo, J. Phys.

Chem. C 2011, 115, 13322-13328.

[42] A. Turunen, M. Linnolahti, V. A. Karttunen, T. A. Pakkanen, P. Denifl, T. Leinonen, J. Mol. Catal. A Chem. 2011, 334, 103-107.

[43] M. D'Amore, R. Credendino, P. H. M. Budzelaar, M. Causa, V. Busico, J.

Catal. 2012, 286, 103-110.

[44] F. Capone, L. Rongo, M. D'Amore, P. H. M. Budzelaar, V. Busico, J.

Phys. Chem. C 2013, 117, 24345-24353.

[45] E. Morra, E. Giamello, S. Van Doorslaer, G. Antinucci, M. D'Amore, V. Busico, M. Chiesa, Angewandte Chemie - International Edition 2015. [46] A. Correa, R. Credendino, J. T. M. Pater, G. Morini, L. Cavallo,

Macromolecules 2012, 45, 3695-3701.

[47] R. Credendino, J. T. M. Pater, D. Liguori, G. Morini, L. Cavallo, J. Phys.

Chem. C 2012, 116, 22980−22986.

[48] R. Credendino, D. Liguori, G. Morini, L. Cavallo, Journal of Physical

Chemistry C 2014, 118, 8050-8058.

[49] R. Credendino, D. Liguori, G. Morini, L. Cavallo, J. Phys. Chem. C 2014,

118, 8050-8058.

[50] R. Credendino, D. Liguori, Z. Fan, G. Morini, L. Cavallo, ACS Catalysis

2015, 5, 5431-5435.

[51] A. Andoni, J. C. Chadwick, S. Milani, H. Niemantsverdriet, P. C. Thune, J.

Catal. 2007, 247, 129-136.

[52] A. Andoni, J. C. Chadwick, H. J. W. Niemantsverdriet, P. C. Thune, J.

Catal. 2008, 257, 81-86.

[53] M. D’Amore, K. S. Thushara, A. Piovano, M. Causà, S. Bordiga, E. Groppo, ACS Catal. 2016, 6, 5786-5796.

[54] K. S. Thushara, E. S. Gnanakumar, R. Mathew, R. K. Jha, T. G. Ajithkumar, P. R. Rajamohanan, K. Sarma, S. Padmanabhan, S. Bhaduri, C. S. Gopinath, J.Phys.Chem.C 2011, 115, 1952-1960. [55] E. S. Gnanakumar, R. R. Gowda, S. Kunjir, T. G. Ajithkumar, P. R.

Rajamohanan, D. Chakraborty, C. S. Gopinath, ACS Catal. 2013, 3, 303-311.

[56] P. Sozzani, S. Bracco, A. Comotti, R. Simonutti, I. Camurati, J. Am.

Chem. Soc. 2003, 125, 12881-12893.

[57] F. Auriemma, C. De Rosa, Chem. Mater. 2007, 19, 5803-5805. [58] F. Auriemma, C. De Rosa, J. Appl. Crystallog. 2008, 41, 68-82. [59] F. Malizia, A. Fait, G. Cruciani, Chem. Eur. J. 2011, 17, 13892-13897. [60] R. Huang, F. Malizia, G. Pennini, C. E. Koning, J. C. Chadwick,

Macromol. Rapid Commun. 2008, 29, 1732-1738.

[61] E. Grau, A. Lesage, S. Norsic, C. Copéret, V. Monteil, P. Sautet, ACS

Catalysis 2013, 3, 52-56.

[62] V. D'Anna, S. Norsic, D. Gajan, K. Sanders, A. J. Pell, A. Lesage, V. Monteil, C. Copéret, G. Pintacuda, P. Sautet, J. Phys. Chem. C 2016,

120, 18075-18087.

(7)

[64] O. Haida, H. Suga, S. Seki, J. Chem. Thermodynamics 1977, 9,

1133-1148.

[65] J. H. S. Geen, Revision of the values of the Heat of formation of normal

alcohols, in; Chem. Ind.: London, 1960, p. 1215-1216.

[66] R. Dovesi, R. Orlando, A. Erba, C. M. Zicovich-Wilson, B. Civalleri, S. Casassa, L. Maschio, M. Ferrabone, M. De La Pierre, P. D'Arco, Y. Noël, M. Causà, M. Rérat, B. Kirtman, Int. J. Quantum Chem. 2014,

114, 1287-1317.

[67] S. J. Grimme, Comput. Chem. 2006, 27, 1787-1799.

[68] B. Civalleri, C. M. Zicovich-Wilson, L. Valenzano, P. Ugliengo,

CrystEngComm 2008, 10, 405−410.

[69] M. D’Amore, R. Credendino, P. H. M. Budzelaar, M. Causá, V. Busico, J.

Catal. 2012, 286, 103-110.

(8)

Calculations at DFT-D level, coupled with FT-IR spectroscopy of CO adsorbed at 100 K allows predicting and validating the morphology of disordered MgCl2

nano-crystals formed in the presence of different alcohols as Lewis bases.

K. S. Thushara, Maddalena D’Amore,* Alessandro Piovano, Silvia Bordiga and Elena Groppo*

Page No. – Page No.

The Influence of Alcohols in Driving the Morphology of MgCl2

Riferimenti

Documenti correlati

3.6 Validation of the multilinear regression model with non-polar solvation, Lennard-Jones and electrostatic contributions scaled 95 3.7 Experimental and calculated folding free

itpcm newsletter www.itpcm.sssup.it December 2009 by Carolina De Simone* IRAN STATE-AID REFORM CONTROLLING THE CONSENSUS *Graduate Student, Scuola Superiore Sant’Anna.. Scholars

Tuttavia non la Matilde storica, così come ce la tramandano le scarne informazioni degli annali, ma una Matilde trasfigurata, mitizzata, attraverso contatti – talvolta indi- retti

Here, to make up for the relative sparseness of weather and hydrological data, or malfunctioning at the highest altitudes, we complemented ground data using series of remote sensing

The face-to-face kick-off event (2 days) was based on a Master Class (or lecture) approach (properly leveled to a practitioner level), aimed to present the course program, giving

Equilibrium phase diagram for aqueous EtOH solutions of CdCl 2 -bpp system at a 20 mM total concentration of the reactants, with indication of the stability regions for 3D-sra,

Milanese F., Don Milani - Quel priore seppellito a Barbiana, Firenze, Libreria editrice fiorentina, 1987.. Mocciaro R., I cattolici contro la scuola confessionale, Bologna, La

This result strongly suggests that we are observing discrete shifts from part-time to full-time work, as conjectured by Zabalza et al 1980 and Baker and Benjamin 1999, rather