• Non ci sono risultati.

Continuous Wave and Pulse EPR Characterization of Open-Shell Ti3+ Ions Generated in Hybrid SiO2–TiO2 Monoliths

N/A
N/A
Protected

Academic year: 2021

Condividi "Continuous Wave and Pulse EPR Characterization of Open-Shell Ti3+ Ions Generated in Hybrid SiO2–TiO2 Monoliths"

Copied!
9
0
0

Testo completo

(1)

Continuous Wave and Pulse EPR Characterization of Open-shell Ti

3+

Ions Generated

in Hybrid SiO

2

-TiO

2

Monoliths.

Elena Morra1, Andriy Budnyk2, Alessandro Damin1, Mario Chiesa1*

1) Dipartimento di Chimica & NIS Centre, Università di Torino, Via P. Giuria 7, I-10125 Torino, Italy 2) International Research Center "Smart materials", Southern Federal University, 344090, 5, Zorge St.,

Rostov-on-Don, Russian Federation. e-mail: mario.chiesa@unito.it

Key words: EPR spectroscopy, open-shell, porous materials, TS-1 Introduction

Understanding the electronic, chemical and structural properties of transition metal ions (TMI) incorporated in micro and mesoporous frameworks, their interaction with the surrounding matrix and their changes during a particular process is key for the development of new catalytic systems and a shift from deductive to predictive catalysis. Key to the catalytic importance of TMI is the presence of partially filled d shells, leading, when stabilized at solid surfaces, to species with unusual oxidation states and coordination numbers.1 Frequently, in particular in the case of first row

transition metals, open-shell (radical type) species are encountered, either as active species or as intermediates within the catalytic cycle. Elucidating the role and influence of such species, which can be of paramount importance in promoting selective radical-type processes as demonstrated in a variety of metallo-enzymes, is presently one of the most challenging endeavours both from an experimental and theoretical perspective. In this context, improvements in ex-situ and in-situ characterization methodologies for the establishment of structure-function relationships is clearly of utmost importance. Where open-shell species are involved, Electron Paramagnetic Resonance (EPR)2,3,4 – an advanced spectroscopic technique akin to Nuclear Magnetic Resonance (NMR) - is

without doubt the technique of choice for investigating and fully interrogating the reaction system. The introduction of pulsed and high-field EPR techniques has boosted the sensitivity of the technique in the past years.5,6,7,8 Nevertheless, this technique is still largely underexploited in the

field of catalysis. In this paper we wish to illustrate a methodological approach involving the use of advanced EPR methods for the characterization of paramagnetic transition metals in heterogeneous catalysis. To do so we have chosen a prototypical example consisting of Ti ions inserted in a mesoporous silica framework obtained via sol-gel method,9 a system that bears similarity to the

(2)

selective oxidation of hydrocarbons.10,11 The distinctive feature of titanium incorporated in

silica-based structures is the possibility of isolating Ti ions with unusual tetrahedral coordination. It is in fact this open coordination that primes the isolated metal centers for the remarkable inner-sphere redox reactivity of TS-1. An interesting, but rarely addressed question, is related to the redox activity of framework Ti4+ ions and in particular on the structural details of reduced Ti3+ species.

The generation and characterization of tetrahedral Ti3+ species in open frame materials is relatively

unexplored, although the importance of such an open shell configuration in catalysis is well recognized. The reactivity of open shell Ti3+ compounds in catalysis is at the heart of important

processes such as olefins polymerization in Ziegler-Natta catalysis, the reactivity of Ti3+-MOF

materials12 and is of general interest in the context of designing earth abundant based transition

metal catalyst, capable of promoting selective radical-type transformations through the participation of the Singly Occupied Molecular Orbital (SOMO) to the reaction. Moreover, since EPR spectra are very sensitive to local coordination symmetry, reduced Ti species can be used as local probes for the coordination environment of incorporated Ti ions in different matrixes.

In our group we have devoted significant efforts in the characterization of the properties of Ti3+

framework centers in several open framework systems including TS-1,13 TiAlPO 14 and ETS-10 15.

Here we add a new system where tetrahedrally coordinated framework Ti4+ ions are reduced with

alkyl aluminium compounds. Alkyl aluminium compounds are indispensable activating agents for olefin polymerization with Ziegler-Natta catalysts, where formation of the active sites is generally accepted to proceed by alkylation and reduction of the titanium species in the catalyst during the activation reaction. Alkyl aluminium compounds are also prototypical precursors to the industrially important methyl aluminoxane co-catalysts for olefin polymerization and have been used to introduce well-defined extra framework aluminium ions in zeolite materials. 16

Here, as a complement to previous studies, we use triethylaluminium (TEA) vapors to reduce framework isolated Ti centers and explore their structure by means of continuous wave (CW) and 2D pulse EPR techniques revealing the local coordination environment through detection of hyperfine interactions of the unpaired electron localized on the Ti3+ ions with magnetically active

nuclei of the matrix, including natural abundant framework 29Si.

Theoretical background. EPR spectra of tetrahedrally coordinated Ti3ions.

In the case of titanium ions, the dominating interactions are the electron Zeeman interaction (characterized by the g matrix) and the hyperfine interactions of the unpaired electron with the isotopes 47Ti (I=5/2, natural abundance 7.75%) and 49Ti (I=7/2, natural abundance 5.51%). Due to

(3)

from the CW EPR powder spectrum is usually limited to the electron Zeeman interaction, which is a useful reporter of the symmetry of the site.17,18

In the classical crystal-field theory, the EPR parameters are explained by using the perturbation formulas based on the one-spin-orbit (SO)-parameter model. Under the effect of a tetrahedral crystal field the degenerate five d orbitals are split to give a pair and a triplet of degenerate levels separated by an energy difference ΔTd. Considering a tetragonal distortion axis along the z-direction,

the g values are given, to first order, by the following equations in the case of tetragonal compression and elongation, respectively:

g∥≅ge and g⊥≅ge-6λ/ΔTd (1)

g≅ge-8λ/ΔTd and g≅ge-2λ/ΔTd (2)

In Eq.1 and Eq.2  is the spin–orbit coupling constant (154 cm-1 for Ti3+) and Δ

Td is the energy

separation between the degenerate triplet and doublet levels in the cubic tetrahedral field. Thus, for a tetragonal compression g < g∥  ge is expected, while for a tetragonal elongation g < g < ge.

2D Pulse EPR methods, like HYSCORE, allow the identification and investigation of the hyperfine and nuclear quadrupole interactions of the magnetic ligand nuclei (e.g. 1H, 29Si, 13C, …).

HYSCORE spectroscopy is a two-dimensional (2D) experiment in which a mixing pulse creates correlations between the nuclear frequencies in two different electron spin (mS) manifolds. For a

system with an unpaired electron (S=1/2) interacting with a nucleus with spin I=1/2 (e.g. 1H and 29Si), the spin Hamiltonian can be described in terms of the g matrix and the hyperfine matrix A. In

general, the hyperfine matrix elements are the sum of an isotropic hyperfine coupling, aiso, and a

point–dipolar contribution, which depends on the distance r between the unpaired electron and the nucleus.

For distances r between the electron and nuclear spin greater than approximately 0.25 nm, the anisotropic part of the hyperfine interaction can be used to calculate the electron–nuclear distance and orientation with the electron–nuclear point–dipole formula:

T = μ0

4 π ℏgeβegnβn

k≠ N

ρk(3~nknk−1)

rk3 (3)

where the sum is over all nuclei with spin population k at distance rk from the nucleus with the

electron–nucleus unit vector nk. For an axial interaction with positive gn, T = [–T, –T, 2T].

For an S=1/2, I=1/2 system, the nuclear frequencies in the two mS manifolds are given by:

ν

α , β

=

[(

A

2

±

ν

I

)

2

+(

B

2

)

2

]

(4)

(4)

where νI is the nuclear Zeeman frequency and A and B describe the secular and pseudosecular part of the hyperfine coupling and are related to aiso and T.19 In the HYSCORE experiment, the

correlations between the nuclear frequencies lead to the cross-peaks (

ν

α ,

ν

β ) and (

ν

β ,

ν

α )

in the 2D plot. The interpretation of HYSCORE spectra of disordered S=1/2, I=1/2 systems has been discussed by several authors20

Results and discussion

The calcined sample is EPR silent as expected for Ti species in the form of diamagnetic Ti4+. Upon

contacting the solid with TEA vapors an intense EPR spectrum with pseudo-axial symmetry is observed (Figure 1a), similar to the one observed in the case of TS-1. The simulation analysis shows that the spectral profile is dominated by a species with slightly rhombic g tensor (gx = 1.900,

gy = 1.941, gz = 1.983. The g values are close to those observed for Ti3+ in TS-1 (Table 1) and

indicate a similar crystal-field symmetry. Substituting the g experimental values in the above equations,  can be estimated to be of the order of 12500 cm-1, which fits relatively well, given the

inherent approximation, with values reported for d-d transitions of tetrahedral Ti3+.21

System gxx gyy gzz Axx Ayy Azz Ref. Ti-SiO2  0.0051.900  0.0021.941  0.0021.983 29Si (1) -6.2  0.2 -6.6  0.2 -9.5  0.2 This work 29Si (2) -4.3  0.2 -4.3  0.2 -4.9  0.2 29Si (3) 0.2  0.2 0.2  0.2 -2.5  0.2 1H (1) -0.2  0.2 -0.2  0.2 7.5  0.4 1H (2) -0.1  0.1 -0.1  0.1 0.5  0.2 TS-1  0.0011.922  0.0011.939  0.00021.9897 29Si (1) -6.8  0.2 -7.1  0.2 -10  0.2 13 29Si (2) -0.3  0.3 -0.3  0.3 -1.6  0.4 1H -0.5  0.2 -0.5  0.2 2.5  0.5

Table 1. g tensor and 29Si and 1H hyperfine coupling constants of Ti-SiO

2 and TS-1. Hyperfine coupling values are

(5)

Figure 1. Experimental (black) and computer simulated (red) a) ESE-detected EPR and b) CW-EPR spectra of Ti/SiO2

reduced with TEA.

The g factor analysis of the CW-EPR spectrum thus strongly indicates the formation of framework Ti3+ ions in tetrahedral coordination upon reaction of the dehydrated sample with TEA vapors. In

order to confirm this analysis and to ascertain the local coordination environment of the formed Ti3+

species, HYSCORE experiments were performed. Figure 2 shows the HYSCORE spectra of the reduced sample taken at two magnetic field settings corresponding to the turning point of the EPR spectrum (see arrows in Figure 1). The spectrum is characterized by the presence in the (+,+) quadrant of an extended ridge centred at the proton nuclear Larmor frequency. The ridge extension is approximately 8 MHz and demonstrates the presence of a proton in the first coordination sphere of the Ti3+ ion. Simulation analysis of the 1H HYSCORE signal at the two different observer

positions, allows recovering the full hyperfine tensor (Table 1), which can be decomposed in the Fermi contact (aiso=1.5±0.5 MHz) and dipolar components (T=3±0.5 MHz).

(6)

Figure 2. Experimental (up) and simulated (down) HYSCORE spectra of Ti/SiO2 reduced with TEA. The spectra were

recorded at magnetic field positions a) 351.0 mT and b) 358.2 mT. Experimental spectra recorded with τ values of 192 ns and 224 ns were added together after Fourier transform. The cross peaks not simulated in spectrum a) are ascribed to

47,49Ti.

Assuming a pure dipolar through-space interaction, the electron-proton distance can be obtained from equation 5: 0 3 1 4 e n e n T g g r    (5) with r being the distance between the unpaired electron localized in the Ti3+ d orbitals and the 1H

nucleus. Using the experimental value obtained from the simulation, the distance from the Ti3+

center and the proton can be estimated to be of the order of 0.29 nm. We notice that the proton hyperfine coupling for the Ti/SiO2 system is much larger, corresponding to shorter Ti-H distance,

with respect to the TS-1 case, where only protons at a distance of the order of 0.43 nm were observed.12 This result may be explained considering the higher flexibility of the SiO

2 framework

compared to that of TS-1, which allows the charge compensating proton to sit closer to the Ti3+ ion.

In addition to the proton signal, the HYSCORE spectrum features a correlation ridge with maximum extension of approximately 10 MHz, centered at Si in the (+,+) quadrant and a pair of

(7)

assigned to the hyperfine interaction of the unpaired electron with natural abundance 29Si (4.67%)

nuclei of the framework. The spectrum recorded at the g position (Figure 2a) shows extra cross

peaks in both quadrants, which are due to the 47Ti (I=5/2, g

n=−0.31484) and with 49Ti (I=7/2,

gn=−0.31491). These signals are not visible in the spectrum recorded at the other field position,

being the corresponding hyperfine interaction exceedingly large. We notice that this is, to the best of our knowledge, the first HYSCORE observation of the 47,49Ti interaction for Ti3+ species. The full

hyperfine tensor could not be obtained because of the absence of the perpendicular component. Although this component is in principle detectable via CW-EPR, the signals are obscured completely by the intense spectrum of titanium with zero nuclear spin hyperfine couplings. An estimate of the hyperfine coupling was done by means of computer simulation of the experimental HYSCORE spectrum and is of the order of 6 MHz, in line with values observed in beryl single crystals.22 The full set of 29Si hyperfine coupling constant parameters extracted by means of

computer simulation of the experimental spectra (Figure 2) are listed in Table 1, where the signs were chosen based on the point dipolar approximation and considering that gn(29Si) < 0.

The values are in line with those obtained for the TS-1 catalyst and compare nicely with those reported by Zamani et al.23 in the case of VO2+ groups incorporated in mesoporous silica materials.

The HYSCORE spectra, together with the g factors of the CW-EPR spectrum (characteristic of a local tetrahedral symmetry) provide thus unique and direct evidence for the framework incorporation of the formed Ti3+ species and absence of TiO

2 phase segregation. We remark that this

result is non trivial as, although the starting material features framework incorporated Ti4+ ions,

chemical reduction to Ti3+ not necessarily implies that the reduced species preserves its framework

nature and local coordination symmetry. The hyperfine tensors reported in Table 1 can be decomposed in the isotropic (Fermi contact) and anisotropic (dipolar) tensors. The strongly coupled

29Si interaction is dominated by the Fermi contact term a

iso = -7.5 MHz while the second largest

coupling gives a value of -4.5 MHz. The third coupling leading to aiso = -0.7 MHz is associated to

remote, matrix nuclei and we will not comment it further. The origin of this fairly large isotropic interaction can be rationalized considering a spin density transfer through the directly coordinated oxygen ions via a through-bond mechanism. The isotropic constant is determined mainly by the unpaired spin density in the 3s orbital of the silicon atoms and is proportional to the value of a0 =

-3995.33 MHz, which is computed for unit spin density in this orbital.24 Including a correction for

departure of the g value (giso[Ti3+/TS-1] = 1.941) from the free electron value (ge = 2.0023) the spin

(8)

(3 ) 0 iso e Si s iso A g a g   (6) to be approximately 0.20% and 0.1% for the two cases. As discussed elsewhere,12 the amount of

spin density transfer is expected to depend markedly on bond angle and distance of the fragment Ti-O-Si, making the Fermi contact term a sensitive structural probe. Comparison can be set to the case of framework substituted Ti3+ ions in AlPO-5 materials.13 In that case Ti3+ species experience the

same tetrahedral field as testified by the similarity of the g matrix components but the nearest cations are 31P instead of 29Si. Also in that case the 31P hyperfine interaction is dominated by the

isotropic term, leading to a spin density population in the P 3s orbital of  0.20% in line with the present results.

Based on this analysis we can then rationalize the observation of two types of 29Si hyperfine

interactions in terms of framework incorporation of the Ti3+ species, whereby the difference in the

hyperfine couplings can be naturally explained considering the different bonding geometry (length and angle) of the Ti-O-Si fragments. The observation of these 29Si hyperfine interactions provides

direct and unambiguous evidence for incorporation of the transition metal ion into the siliceous framework.

Conclusions

The reaction of triethyl aluminium with hybrid SiO2−TiO2 mesoporous monoliths synthesized via

sol−gel method, was followed by means of CW-EPR and HYSCORE spectroscopies. The results provide evidence for the reducing power of triethyl aluminium towards isolated Ti4+ ions through

the formation of open-shell Ti3+ ions atomically dispersed and fully incorporated in the SiO 2

framework with a local structure similar to that of titanium silicalite and no segregation of TiO2

phases at this level of Ti doping. The observation of well resolved superhyperfine interactions with magnetically active framework ions provides a compelling evidence for framework substitution and allows identifying specific active sites. Interestingly, HYSCORE experiments reveal a much larger proton interaction for this system with respect to TS-1, which may be associated to a higher degree of flexibility of the SiO2 framework, compared to that of TS-1.

The implications of these results are manifold; from one side they provide a simple and effective route to generate accessible tetrahedrally coordinated open-shell Ti3+ sites, which may open new

and unexplored reaction pathways in coordination chemistry and redox catalysis, on the other hand this approach provides a robust framework to be extended to other oxide or halide based systems in the hope to achieve a more general understanding of the surface chemistry of alkylaluminum compounds. Finally these results demonstrate that advanced EPR methods can be a valuable way to

(9)

afford a fundamental understanding of the atomic scale structure of paramagnetic species in heterogeneous catalysts.

Materials and Methods

Sample preparation

The Ti/SiO2 sample (Ti loading was 2 wt%) was prepared by the sol-gel method described in ref 9

using titanium isopropoxide (TTIP) as Ti source. The amount of TTIP was chosen to have Ti/Si molar ratio of 2 M%.The sample was dehydrated at 923 K in dynamic vacuum to a final pressure of 10-5 mbar and then calcined at 723 K in O

2 atmosphere in order to remove organic residues. The

dehydrated sample was reacted for 6 minutes with triethylaluminum (TEA) (Aldrich) vapors generated in-situ employing the setup described elsewhere.12 Before use, TEA was purified by

freeze-pump-thaw method. Spectroscopic characterization

X-band CW EPR spectra were recorded at T = 77 K on a Bruker EMX spectrometer (mw frequency 9.46 GHz) equipped with a cylindrical cavity. A microwave power of 0.1 mW, a modulation amplitude of 0.08 mT and a modulation frequency of 100 kHz were used.

X-band pulse EPR experiments were performed at T = 15 K on a Bruker ELEXYS 580 EPR spectrometer (mw frequency 9.76 GHz) equipped with a liquid-helium cryostat from Oxford Inc. The magnetic field was measured by means of a Bruker ER035M NMR gaussmeter.

Electron-spin-echo (ESE) detected EPR experiments were performed with the pulse sequence π/2−τ−π−τ−echo. Pulse lengths tπ/2 = 16 ns and tπ = 32 ns, a τ value of 200 ns and a 0.5 kHz shot

repetition rate were used.

Hyperfine Sublevel Correlation (HYSCORE)25 experiments were carried out with the pulse

sequence π/2−τ−π/2−t1−π−t2−π/2−τ−echo. A four-step phase cycle was applied for eliminating

unwanted echoes. Pulse lengths tπ/2 = 16 ns and tπ = 32 ns, and a 0.5 kHz shot repetition rate were

used. The time intervals t1 and t2 were varied in steps of 16 ns starting from 96 ns, generating a [300

x 300] matrix. The time traces of the HYSCORE spectra were baseline corrected with a third-order polynomial, apodized with a Hamming window, and zero-filled. After two-dimensional Fourier transformation, the absolute-value spectra were calculated. Spectra recorded with different τ values specified in the figure captions were added to eliminate blind-spot effects.

All of the EPR spectra were simulated with the Easyspin package.26

(10)

1 Dyrek K, Che M (1997) Chem Rev 97:305-332.

2 Chiesa M, Giamello E, Che M (2010) Chem Rev 110:1320-1347. 3 Morra E, Giamello E, Chiesa M (2017) J Magn Reson 280:89-102. 4 Van Doorslaer S, Murphy DM (2012) Top Curr Chem 321:1-39.

5 Telser J, Krzystek J, Ozarowski A (2014) J Biol Inorg Chem 19:297-318. 6 Van Doorslaer S (2017) J Magn Reson 280:79-88.

7 Jeschke G (2016) eMagRes 5:1459-1475.

8 Spindler PE, Schöps P, Bowen AM, Endeward B, Prisner TF (2016) eMagRes 5:1477-1492. 9 Budnyk A, Damin A, Bordiga S, Zecchina A (2012) J Phys Chem C 116:10064−10072. 10 Notari B (1996) Adv Catal 41:253-334.

11 Bordiga S, Bonino F, Damin A, Lamberti C (2007) Phys Chem Chem Phys 9:4854-4878. 12Brozek, C. K,, Dincă M. (2013) J. Am. Chem. Soc., 135, 12886-12891.

13 Morra E, Giamello E, Chiesa M (2014) Chem Eur J 20:7381–7388.

14 Maurelli S, Vishnuvarthan M, Chiesa M, Berlier G, Van Doorslaer S (2011) J Am Chem Soc 133:7340– 7343.

15 Morra E, Maurelli S, Chiesa M, Giamello E (2015) Top Catal 58:783–795.

16 Kerber RN, Kermagoret A, Callens E, Florian P, Massiot D, Lesage A, Copéret C, Delbecq F, Rozanska

X, Sautet P (2012) J Am Chem Soc 134:6767-6775.

17 Figgis BN (1967) In: John Introduction to ligand fields, Wiley & Sons, New York.

18 Weil JA, Bolton JR, Wertz JE (1994) In: Electron paramagnetic resonance: elementary theory and practical applications, John Wiley, New York.

19 Schweiger A, Jeschke G (2001) In: Principles of electron paramagnetic resonance, Oxford University

Press, Oxford.

20 Pöppl A, Kevan L (1996) J Phys Chem 100:3387-3394.

21 Griffith JS (1964) In: The theory of transition-metal ions, Cambridge University Press, Cambridge.

22 Solntsev VP, Yurkin AM (2000) Cryst Reports 45:128-132.

23 Zamani S, Meynen V, Hanu AM, Mertens M, Popovici E, Van Doorslaer S, Cool P (2009) Phys Chem

Chem Phys 11:5823-5832.

24 Fitzpatrick JAJ, Manby FR, Western CM (2005) J Chem Phys 122:084312-1-12. 25 Höfer P, Grupp A, Nebenfür H, Mehring M (1986) Chem Phys Lett 132:279−282.

Riferimenti

Documenti correlati

Si noterà come in questa lista figurino nomi ormai familiari (Erzuah, Yanzu Aka e Kwasi) a segnalare in maniera inequivocabile il tentativo da parte dei soggetti

oral como ferramenta de análise com mulheres agricultoras inseridas no Movimento de Mulheres Trabalhadoras Rurais na região Noroeste do Rio Grande do Sul, procuramos discorrer

By developing and implementing effective tools and procedures, we were able to convert CoBiS datasets, composed of different data formats, into RDF, to interlink them with

L’interesse per la ricerca teorica sulle forme narrative fondata su principi e procedimenti di indagine innovativi si manifesta in Rus- sia, com’è noto, a partire dalla seconda

We presently consider the task of automatically assessing the conceptual similarity (that is, given a pair of concepts the problem is to pro- vide a score that expresses in how far

Statement Winemaking Marketing Get Higher Quality Wines Simplify the Work in the Cellar Produce More Natural Wines Use a Yeast Permanently Residing in a Certain Terroir Get

A clin- ical and instrumental diagnosis was performed much more than half by Pediatricians of Northern and in all Central Italy than pediatricians living in Southern of Italy (just

Finché la relazione tra finito e infinito non è vista come questa relazione inseparabile dell’esser per sé e dell’essere per altro, in cui l’essere per altro è un