Carlos O.
Arean
*a,
Montserrat R.
Delgado
a,
Petr
Nachtigall
b,
Ho Viet
Thang
b,
Miroslav
Rubeš
b,
Roman
Bulánek
c and
Pavla
Chlubná-Eliášová
d
aDepartment of Chemistry, University of the Balearic Islands, 07122 Palma de Mallorca, Spain. E-mail: co.arean@uib.es; Fax: +34 971173426; Tel: +34 971173251
bDepartment of Physical and Macromolecular Chemistry, Faculty of Science, Charles University of Prague, Hlavova 2030, Prague 2, 128 00, Czech Republic
cDepartment of Physical Chemistry, Faculty of Chemical Technology, University of Pardubice, 532 10 Pardubice, Czech Republic
dJ. Heyrovsky Institute of Physical Chemistry, Academy of Sciences of the Czech Republic, 182 23 Prague, Czech Republic
First published on 30th January 2014
Brønsted-acid zeolites are currently being used as catalysts in a wide range of technological processes, spanning from the petrochemical industry to biomass upgrade, methanol to olefin conversion and the production of fine chemicals. For most of the involved chemical processes, acid strength is a key factor determining catalytic performance, and hence there is a need to evaluate it correctly. Based on simplicity, the magnitude of the red shift of the O–H stretching frequency, Δν(OH), when the Brønsted-acid hydroxyl group of protonic zeolites interacts with an adsorbed weak base (such as carbon monoxide or dinitrogen) is frequently used for ranking acid strength. Nevertheless, the enthalpy change, ΔH0, involved in that hydrogen-bonding interaction should be a better indicator; and in fact Δν(OH) and ΔH0 are often found to correlate among themselves, but, as shown herein, that is not always the case. We report on experimental determination of the interaction (at a low temperature) of carbon monoxide and dinitrogen with the protonic zeolites H-MCM-22 and H-MCM-56 (which have the MWW structure type) showing that the standard enthalpy of formation of OH⋯CO and OH⋯NN hydrogen-bonded complexes is distinctively smaller than the corresponding values reported in the literature for H-ZSM-5 and H-FER, and yet the corresponding Δν(OH) values are significantly larger for the zeolites having the MWW structure type (DFT calculations are also shown for H-MCM-22). These rather unexpected results should alert the reader to the risk of using the O–H frequency shift probed by an adsorbed weak base as a general indicator for ranking zeolite Brønsted acidity.
At variance with aqueous acid solutions for which the corresponding pKa provides a quantitative measure of acid strength, no clear-cut measurements have yet been found for solid acids. Initial attempts at measuring acid strength of protonic zeolites by using amine-based Hammett indicators did not yield consistent results,13,14 and indeed some objections to the use of that method were raised.15,16 Currently, several instrumental techniques are used to gather increasing information on the (relative) Brønsted acidity of protonic zeolites. Among them, calorimetry of an adsorbed base, temperature programmed desorption, IR spectroscopy and solid state NMR17–28 are sometimes combined with catalytic tests of acid strength.
Due (in part) to the ease of its usage, IR spectroscopy of an adsorbed weak base is (arguably) the most widely used technique nowadays for measuring Brønsted acid strength of protonic zeolites, carbon monoxide being the preferred base molecule, although dinitrogen could equally well be used.29–34 In principle, proton transfer between an acid zeolite, ZOH, and a sufficiently strong (adsorbed) base, B, would involve both, hydrogen-bonded and ion-paired intermediates, following eqn (1) below.
ZOH + B ⇆ ZOH⋯B ⇆ ZO−⋯HB+ ⇆ ZO− + BH+ | (1) |
However, for weak bases the process stops before the ion pair is formed, yielding only the hydrogen-bonded species, ZOH⋯B, which can easily be monitored by IR spectroscopy because hydrogen bonding brings about a distinctive bathochromic shift, Δν(OH), of the O–H stretching mode of the zeolite Brønsted-acid hydroxyl group. Moreover, one would expect that the stronger the Brønsted acid site, the larger should be the value of Δν(OH) (for any given weak base). Hence, direct measurement of Δν(OH) in the infrared spectrum of the hydrogen-bonded complex could, in principle, afford ranking of zeolite acidity. This method was pioneered some time ago by Paukshtis and Yurchenko,35 and more recently reviewed in some detail by Dwyer et al.36 and by Rigby et al.,37 who confirmed the previously proposed logarithmic relationship between Δν(OH) and zeolite acid strength, as probed by adsorbed CO or dinitrogen. On account of its simplicity, direct correlation of Δν(OH) with acid strength is broadly invoked to rank zeolite acidity. However, this method is not free from a number of possible pitfalls, as pointed out by several authors.38–43
Recently, we found out a notable discrepancy in the apparent Brønsted acidity of H-MCM-22 (a zeolite belonging to the MWW structure type) when results obtained by (i) the Δν(OH) method and (ii) direct measurement of the enthalpy of hydrogen-bonding formation with adsorbed CO and dinitrogen were compared with those reported in the literature for other protonic zeolites.44 Herein, we revise those experimental results, and compare them with corresponding calorimetric measurements and DFT calculations. In addition, we report experimental measurements on another H-MCM-22 sample having a different Si:Al ratio, as well as on H-MCM-56, which is another zeolite having the MWW structure type. The whole set of experimental results is discussed in the broader context of corresponding available data for other protonic zeolites having different structure types, with a particular reference to H-FER.
It is well known that for both carbon monoxide and dinitrogen, adsorption at a low temperature in protonic zeolites results in hydrogen bonding because of localized interaction of the gas molecules with the Brønsted-acid [Si(OH)Al] groups of the zeolite. Such a process results in progressive erosion of the characteristic O–H stretching band, which is red-shifted when forming the corresponding OH⋯CO or OH⋯N2 complex; simultaneously, the characteristic IR absorption band of the hydrogen-bonded molecule builds up. A set of IR spectra obtained over a temperature range while simultaneously recording IR absorbance, temperature and gas equilibrium pressure inside a closed IR cell can be used to determine the standard enthalpy, ΔH0, and entropy, ΔS0, of the (localized) gas adsorption process.50 The method, termed VTIR spectroscopy, was explained in detail elsewhere,51 and tested not only for hydrogen bonding but also for studying the thermodynamics of gas adsorption on a large number of alkaline zeolites.52,53 However, in order to facilitate understanding of the results reported here, an abridged account is given below.
Referring to hydrogen bonding, let eqn (2) below represent the adsorption equilibrium of a molecule, M, on the adsorption centre, ZOH, representing a zeolite Brønsted-acid site:
ZOH(s) + M(g) ⇆ ZOH⋯M(ads) | (2) |
θ = K(T)p/[1 + K(T)p] | (3) |
K(T) = exp(−ΔH0/RT)exp(ΔS0/R) | (4) |
ln{θ/[(1 − θ)p]} = (−ΔH0/RT) + (ΔS0/R) | (5) |
ln{A/[(AM − A)p]} = (−ΔH0/RT) + (ΔS0/R) | (6) |
All calculations were performed with the periodic model using the VASP program package.56,57 The Perdew–Burke–Ernzerhof (PBE) exchange–correlation functional,58,59 the projector augmented wave approximation (PAW)60,61 and the plane wave basis set with a kinetic energy cut-off of 400 eV were used; the Brillouin-zone sampling was restricted to the Γ-point. Geometry optimizations were performed with fixed UC volume and shape, while relaxing the positions of all atoms. At this PBE-optimized geometry the interaction energies were also calculated with dispersion-corrected DFT methods, PBE-D2,62 PBE-D3,63 and vdW-DF2,64 in particular. The C–O stretching frequency of carbon monoxide adsorbed on Brønsted acid sites and the O–H stretching frequencies were evaluated using the ωCO/rCO and νOH/rOH correlation schemes, respectively, described in detail in ref. 54. Only the PBE exchange–correlation functional was used for calculations of vibrational frequencies within the ω/r correlation scheme, since this correlation is with respect to an accurate coupled cluster, CCSD(T), the level of theory following the method previously established.65
Zero-point energies (ZPE) were calculated within the harmonic approximation considering 6 degrees of freedom for CO and 6 degrees of freedom for the Brønsted-acid OH group for the case of Al atoms in the T1 position. Two displacement steps in each direction, with a step size of 0.005 Å, were used to calculate numerically the corresponding second derivatives. Our previous investigation of CO interaction with H-FER54 justifies the use of this ZPE correction for all of the Brønsted-acid sites investigated herein. Adsorption enthalpies are reported for 0 K only.
The substitution energy, ΔE(Al,H), was evaluated following the same strategy as that used by Li et al.66 Therefore, substitution energies calculated herein within the periodic DFT model can be directly compared with those obtained previously based on cluster models.66
To verify the accuracy of various density functional theory-type methods, accurate CCSD(T) calculations were performed on the cluster model shown in Fig. 2. Interaction energies were calculated for CO interacting primarily with the Brønsted-acid site (represented by the 2T cluster model) and simultaneously with a model of zeolite framework represented by a 1T cluster model; CS symmetry constrain was applied and CO was fixed along the OB(2T) and Si(1T) axis as indicated in Fig. 2. Interaction energies are plotted as a function of separation, RO(2T)–Si(1T), between the two cluster models. Reported CCSD(T) interaction potentials are extrapolated to the complete basis set limit, DFT interaction potentials were obtained using the quadruple-ζ basis set with polarization functions.67,68
Al position | Brønsted site | Interaction energy | |||
---|---|---|---|---|---|
PBE | PBE-D2 | PBE-D3 | vdW-DF2 | ||
T1 | O3 | −28.0 | −36.1 | −34.9 | −31.7 |
T2 | O10 | −28.1 | −42.1 | −43.8 | −42.3 |
T3 | O13 | −26.8 | −41.4 | −42.2 | −39.0 |
T4 | O3 | −28.6 | −36.8 | −35.5 | −32.6 |
T5 | O8 | −20.7 | −34.2 | −38.9 | −40.5 |
T6 | O2 | −23.2 | −35.8 | −35.3 | −33.5 |
T7 | O4 | −3.5 | −13.2 | −14.3 | −18.1 |
T8 | O10 | −28.6 | −42.6 | −44.4 | −43.4 |
Results reported in Table 1 are truly disappointing in the light of the experimentally determined CO adsorption enthalpy value of ΔH0 = −22.5 kJ mol−1 (see below). Even when the ZPE correction is accounted for (less than 4 kJ mol−1) adsorption enthalpies calculated with dispersion-corrected methods are about 50% overestimated with respect to reliable experimental data. Even the results obtained at the PBE level (not accounting for dispersion interactions) appear to be overestimated with respect to experiment.
To understand the discrepancy between calculated and experimental results, accurate CCSD(T) calculations were carried out for the cluster model that mimics the simultaneous interaction of CO with a Brønsted-acid site and with the zeolite framework at the opposite side of the channel or cavity (Fig. 2). Results obtained at the PBE level for large RO(2T)–Si(1T) indicate that the local interaction of CO with just the Brønsted-acid site is overestimated by 2.8 kJ mol−1, and this overestimation becomes significantly smaller (about 1 kJ mol−1) for RO(2T)–Si(1T) ≈ 8 Å, which is a typical situation for zeolites. In contrast, results obtained at the PBE-D2 level clearly show that the interaction energy is largely overestimated (6.2–6.7 kJ mol−1) for the whole range of RO(2T)–Si(1T) distance investigated. These results show that the PBE functional gives a more accurate description of this system than dispersion-corrected functionals. Although the relatively good performance of PBE should result from fortuitous error cancellations, its use for describing the CO–H-MWW system seems to be justified. Results reported below were all obtained at the PBE level.
Relevant geometrical parameters describing the Brønsted-acid site (BA site) with and without adsorbed CO, O–H and C–O stretching frequencies, and CO adsorption enthalpy are reported in Table 2 for all eight distinguishable framework Al positions. Only the results obtained for the energetically most stable BA sites in the vicinity of each framework Al are reported in this table; the complete set of results obtained for all non-equivalent BA sites in H-MCM-22 is reported in the ESI† (Table S1). A representative set of BA sites with and without adsorbed CO is depicted in Fig. 3.
Al position | Brønsted site | ΔE(Al,H)a | Geometrical data | Stretching frequencies | ||||||
---|---|---|---|---|---|---|---|---|---|---|
Al–O–Sib | r(OH) | r(OH) + COc | ΔH0 (0 K) | ν(OH) | ν(OH) + COc | Δν(OH) | ν(CO)c | |||
a Relative energies for framework Si substitution with Al/H with respect to the substitution in the T7 position. b Al–O–Si angle (in deg) for the BA site. c For the CO adsorption complex on the BA site. d BA sites involved in intra-zeolite hydrogen bonding. e Not corrected for ZPE. | ||||||||||
T1 | O3 | 17 | 128.1 | 0.9755 | 1.0020 | −24.4 | 3629 | 3280 | 349 | 2183 |
T2 | O10 | 18 | 129.2 | 0.9756 | 0.9994 | −24.5 | 3628 | 3314 | 314 | 2181 |
T3 | O13 | 21 | 132.4 | 0.9765 | 1.0022 | −23.2 | 3617 | 3278 | 339 | 2179 |
T4 | O3 | 20 | 129.0 | 0.9759 | 1.0031 | −25.0 | 3624 | 3266 | 358 | 2183 |
T5d | O8 | 12 | 138.4 | 0.9919 | 0.9994 | −17.1 | 3414 | 3315 | 100 | 2178 |
T6 | O2 | 13 | 127.7 | 0.9761 | 0.9968 | −19.6 | 3621 | 3349 | 272 | 2177 |
T7d | O4 | 0 | 128.2 | 0.9919 | 0.9923 | −3.5e | 3413 | 3408 | 5 | 2139 |
T8 | O10 | 24 | 131.6 | 0.9753 | 0.9985 | −25.0 | 3633 | 3326 | 306 | 2180 |
Fig. 4 (A) Representative variable-temperature IR spectra (O–H stretching region) of CO adsorbed on H-MCM-22 (Si:Al = 24.5:1). The spectra are shown in the difference mode (zeolite blank subtracted). Temperature (in K) and equilibrium pressure (mbar, in brackets) as follows: 1, 154 (6.52); 2, 160 (6.99); 3, 166 (7.37); 4, 172 (7.72); 5, 178 (8.01); 6, 184 (8.29); 7, 190 (8.55); 8, 196 (8.75); 9, 202 (8.93); 10, 208 (9.11); 11, 214 (9.24). The inset shows the blank zeolite spectrum (black) and the effect of dosing with CO at 77 K (grey). (B) Representative variable-temperature IR spectra (O–H stretching region) of CO adsorbed on H-FER (Si:Al = 27.5:1). The spectra are shown in the difference mode (zeolite blank subtracted). From 1 to 9, temperature goes from 167 to 224 K; and the equilibrium pressure from 0.57 to 1.75 mbar. The inset shows the corresponding van't Hoff plot (ref. 52). |
Fig. 5 Plot of the left-hand side of eqn (5) against the reciprocal of the temperature for CO adsorbed on H-MCM-22 (Si:Al = 24.5:1 and 16.4:1) and H-MCM-56; data obtained from the O–H stretching bands. R, linear regression coefficient; SD, standard deviation. |
A comment on how band intensity was measured seems pertinent here. Several authors have pointed out that the clearly dissymmetric (and rather broad) O–H stretching band assigned to the Brønsted acid hydroxyl groups of H-MCM-22 (IR absorption band at 3625 cm−1 in Fig. 4A) can be resolved into two or more components.69–72 And in fact, Garrone et al.69 used computer procedures to simulate the band envelope by adding up three bands peaking at 3628, 3618 and 3585 cm−1; the first two having nearly the same intensity, and a much weaker third one. However, our attempts at band resolution did not produce quantitatively reliable results (because of inherent uncertainty about how the band should be subdivided) regarding calculation of CO adsorption enthalpy. Therefore, for the corresponding van't Hoff plot we decided to use the integrated intensity of the hydroxyl band (at 3625 cm−1) as it appears in the VTIR spectra of Fig. 4A. As a matter of fact, the rather regular shape of this band across the whole series of VTIR spectra, together with the fact that the derived van't Hoff plot (Fig. 5) fits to a straight line, strongly suggest that the molar absorption coefficients of the main OH species adding up to the overall band profile should not differ much from each other.
Fig. 6 depicts the C–O stretching region of the same VTIR spectra shown in Fig. 4A. The set of IR absorption bands seen at 2174 cm−1 corresponds to the C–O stretching mode of carbon monoxide involved in OH⋯CO complexes. The weaker and broader band at about 2138 cm−1, which comes (mainly) from weakly physisorbed (liquid-like CO),28,29 is of no concern here. The inset to Fig. 6 shows the van't Hoff plot obtained by applying eqn (6) to the whole set of spectra recorded (band at 2174 cm−1). From this linear plot, the values of ΔH0 = −23(±2) kJ mol−1 and ΔS0 = −140(±10) J mol−1 K−1 were obtained. As expected, these values practically coincide (within experimental error) with those derived from the O–H stretching band of the same OH⋯CO hydrogen-bonded species; and such a close agreement between both sets of results provides further reliability on the methods used. The average of both sets of results gives the final values of ΔH0 = −22.5(±2) kJ mol−1 and ΔS0 = −135(±10) J mol−1 K−1 for the standard enthalpy and entropy of formation of the hydrogen-bonded CO complexes under study.
Fig. 6 Representative IR spectra (C–O stretching region) of CO adsorbed on H-MCM-22 (Si:Al = 24.5:1). From top to bottom, temperature goes from 154 to 214 K; and equilibrium pressure from 6.52 to 9.24 mbar. The inset shows the corresponding plot of the left-hand side of eqn (6) against the reciprocal of the temperature. R, linear regression coefficient; SD, standard deviation. |
As stated in the Introduction, adsorption calorimetry was also used to measure the enthalpy change in hydrogen-bonding of CO with the Brønsted-acid hydroxyl groups of both, H-MCM-22 (Si:Al = 24.5:1) and H-FER (Si:Al = 27.5:1). CO adsorption isotherms, at 303 K, are shown in Fig. 7A and B, respectively, while Fig. 7C shows the corresponding differential heat of adsorption (in the low coverage range). Note that, in order to test reproducibility, two independent sets of measurements were performed for carbon monoxide adsorption on H-MCM-22. The average of the results thus obtained gives the adsorption heat value of 21(±2) kJ mol−1, which agrees (within experimental error) with the value of ΔH0 = −22.5(±2) kJ mol−1 obtained by VTIR spectroscopy. Regarding H-FER, Fig. 7C shows an adsorption heat of about 26(±2) kJ mol−1, to be compared with the corresponding value of ΔH0 = −28.4(±2) kJ mol−1 previously determined by VTIR spectroscopy.54 It is also noteworthy that (up to the coverage of about 0.2) the CO adsorption heat is (very approximately) independent of coverage for both zeolites.
Fig. 8 (bottom inset) depicts the IR spectrum in the O–H stretching region of the zeolite H-MCM-22 (Si:Al = 16.4:1). The IR absorption bands corresponding to silanols (3750 cm−1) and to Brønsted-acid hydroxyl groups (3625 cm−1) appear at the same wavenumbers as those of the previous H-MCM-22 (Si:Al = 24.5:1) sample. Representative VTIR spectra of adsorbed CO are shown (in the difference mode) in the main body of Fig. 8, whereas the top inset depicts the characteristic C–O stretching mode of the OH⋯CO hydrogen-bonded adsorption complex. From the whole set of the obtained VTIR spectra, the van't Hoff plot shown in Fig. 5 was derived. This linear plot gave the values ΔH0 = −21(±2) kJ mol−1 and ΔS0 = −128(±10) J mol−1 K−1 for the standard adsorption enthalpy and entropy (respectively) of formation of the hydrogen-bonded OH⋯CO species. Note that the ΔH0 value is slightly smaller than that of −22.5 kJ mol−1 obtained for H-MCM-22 (Si:Al = 24.5:1), which could be correlated with the slightly smaller value of Δν(OH): −317 cm−1 for H-MCM-22 (Si:Al = 16.4:1) and −320 cm−1 for H-MCM-22 (Si:Al = 24.5:1). Although it is true that such a small difference in wavenumber can hardly be assessed experimentally, particularly when dealing with IR absorption bands as broad as those shown by the hydrogen-bonded adsorption complexes.
The IR spectra corresponding to the zeolite H-MCM-56 (Si:Al = 16:1) are depicted in Fig. 9. The blank spectrum of the zeolite wafer shows the characteristic silanol and Brønsted-acid hydroxyl group bands at 3749 and 3624 cm−1, respectively (very close to the corresponding values of H-MCM-22). The bathochromic shift of the latter band upon CO adsorption is Δν(OH) = −316 cm−1, as seen in the main body of Fig. 9. From the whole series of VTIR spectra of adsorbed CO (some of which are shown in Fig. 9), the corresponding van't Hoff plot (shown in Fig. 5) was obtained, which gave the values of ΔH0 = −20(±2) kJ mol−1 and ΔS0 = −120(±10) J mol−1 K−1 for the formation of hydrogen-bonded OH⋯CO species in H-MCM-56.
VTIR spectra of adsorbed dinitrogen, which can also be used for testing Brønsted acidity of protonic zeolites, will now be considered. As examples, we show series of such spectra in Fig. 10 (for H-MCM-22 having a Si:Al ratio of 16.4:1) and in Fig. 12 (H-MCM-56, Si:Al = 16:1). Corresponding VTIR spectra of dinitrogen adsorbed on H-MCM-22 (Si:Al = 24.5:1) and H-FER (Si:Al = 27.5:1) were recently reported by Delgado et al.,44 and by Nachtigall et al.,54 respectively.
The bottom inset in Fig. 10 shows the effect of dosing with nitrogen, at 77 K, the H-MCM-22 zeolite wafer. Formation of OH⋯NN hydrogen bonded species results in a bathochromic shift of the IR absorption band of Brønsted-acid hydroxyl groups from 3625 down to 3500 cm−1, that is, Δν(OH) = −125 cm−1. It should be noted, however, that the exact magnitude of Δν(OH) depends slightly on the temperature at which the IR spectrum is recorded (as can be seen in the main body of Fig. 10). Formation of the OH⋯NN adducts results in activation in the IR of the N–N stretching mode, which appears at 2332 cm−1 (top inset in Fig. 10). From the whole series of VTIR spectra that were run, some of which are shown in the main body of Fig. 10, the plot of the left-hand side of eqn (5) against the reciprocal of the temperature depicted in Fig. 11 was obtained. From this linear plot the value of ΔH0 = −13.5(±2) kJ mol−1 was derived for the standard enthalpy of formation of the hydrogen-bonded OH⋯NN species. The corresponding entropy change resulted to be ΔS0 = −103(±10) J mol−1 K−1.
Fig. 11 Plot of the left-hand side of eqn (5) against the reciprocal of the temperature for N2 adsorbed on H-MCM-22 (Si:Al = 16.4:1) and H-MCM-56; data obtained from the corresponding O–H stretching bands. R, linear regression coefficient; SD, standard deviation. |
Finally, representative VTIR spectra of dinitrogen adsorbed on H-MCM-56 are shown in Fig. 12. The IR absorption band at 3624 cm−1 (Brønsted-acid OH groups) is shifted down to 3500 cm−1 upon formation of the hydrogen-bonded species, which results in Δν(OH) = −124 cm−1 (bottom inset). From the integral values of absorbance, taken from the band at 3624 cm−1 in the whole series of VTIR spectra, the van't Hoff plot depicted in Fig. 11 was obtained. The corresponding values of ΔH0 and ΔS0 (for the formation of the hydrogen-bonded species) resulted to be −13(±2) kJ mol−1 and −104(±10) J mol−1 K−1, respectively.
For quantitative analysis of the CO adsorption enthalpy a more reliable level of theory would be required. It should be noted that the dispersion corrected functionals tested herein all failed to describe reliably the interaction of CO with Brønsted-acid sites in H-MCM-22. Careful analysis of DFT results obtained for cluster models showed that the problem of dispersion-corrected exchange–correlation functionals stems not so much from the description of dispersion interactions themselves as from just the local (or semi-local) part of the exchange–correlation functional. Results reported in Fig. 2 show that the PBE functional already overestimates the interaction of CO with the Brønsted acid site, mostly due to a too large electrostatic stabilization. It is then not surprising that dispersion-corrected functionals give even a larger discrepancy with respect to the CCSD(T)/CBS level, since the dispersion contribution further stabilizes the adsorption complex. Hence, the problem of the PBE functional cannot be fixed by adding the dispersion interaction (DFT-D2 or DFT-D3). In contrast, the DFT/CC method, which is based on the correction of DFT errors of any kind (and not just dispersion) was used successfully for the description of similar systems, e.g., CO2 adsorption in zeolites and MOFs;73–75 and it could possibly be used even for the system dealt with herein. The development of DFT/CC correction functions for the interaction of CO with Brønsted-acid sites in zeolites, which is currently under progress, will be hopefully reported in a future article, together with an in-depth analysis of DFT failure to describe such an interaction.
The substitution energy, ΔE(Al,H), calculated with the periodic model (Table 2) can be compared with corresponding values obtained by Li et al.66 using cluster models. The only agreement between the two models is that the most stable position for framework Al is the T7 site. The relative energies for different positions of framework Al obtained with the periodic model are qualitatively different from those obtained with cluster models. In addition, the relative stability of Brønsted-acid sites in the vicinity of a particular framework Al site is also qualitatively different when using periodic and cluster models. That led us to conclude that the use of cluster models for the description of Brønsted-acid sites in zeolites is questionable.
To facilitate discussion of the experimental results, Table 3 summarizes the relevant numerical data reported herein for MWW type zeolites, as well as corresponding data previously reported for other protonic zeolites.34,49,54 In all cases, ΔH0 values were determined by VTIR spectroscopy. Δν(OH) values were all measured as peak-to-peak distance in the corresponding difference IR spectra for CO (or N2) doses approaching a coverage of θ = 1. Much larger doses of the adsorbate should be avoided, particularly for carbon monoxide, because at a low temperature that would result in massive formation of a condensed phase and corresponding perturbation of hydroxyl groups and gas adsorption complexes, which can lead to inconsistent Δν(OH) values. Besides data reported in Table 3, Δν(OH) values for some of the same zeolites (but not corresponding ΔH0 values) were also reported by several authors. Among them, Dwyer et al.36 give the values of Δν(OH)(CO) = −308 cm−1 for CO adsorbed (at θ → 1) in H-ZSM-5, and Δν(OH)(CO) = −273 cm−1 for H-Y; Knözinger et al.27 give, for H-ZSM-5, Δν(OH)(CO) = −317 cm−1 and Δν(OH)(N2) = −121 cm−1; Hadjiivanov et al.76 give, for H-FER, Δν(OH)(CO) = −292 cm−1 and Δν(OH)(N2) = −106 cm−1; and Gil et al.77 give Δν(OH)(CO) = −326 cm −1 for carbon monoxide adsorbed (at 173 K) on H-MCM-22. Although some small differences are observed, these values are consistent with those reported in Table 3.
Zeolite | Structure type | Si/Al ratio | ν OH (cm−1) | −Δν(OH)(CO)a (cm−1) | −ΔH0(CO)b (kJ mol−1) | −Δν(OH)(N2)c (cm−1) | −ΔH0(N2)d (kJ mol−1) | Ref. |
---|---|---|---|---|---|---|---|---|
a Red-shift of the Brønsted-acid OH group upon hydrogen bonding with CO. b Standard enthalpy change in the formation of the OH⋯CO complex. c Red-shift of the Brønsted-acid OH group upon hydrogen bonding with N2. d Standard enthalpy change in the formation of the OH⋯N2 complex. | ||||||||
H-Y | FAU | 5.6 | 3645 | 275 | 25.6 | 98 | 15.7 | 34 and 49 |
H-ZSM-5 | MFI | 30 | 3618 | 303 | 29.4 | 116 | 19.7 | 34 |
H-FER | FER | 27.5 | 3605 | 297 | 28.4 | 110 | 19.1 | 54 |
H-MCM-22 | MWW | 24.5 | 3625 | 320 | 22.5 | 125 | 14.5 | This work |
H-MCM-22 | MWW | 16.4 | 3625 | 317 | 21 | 125 | 13.5 | This work |
H-MCM-56 | MWW | 16 | 3624 | 316 | 20 | 124 | 13 | This work |
Table 3 shows that the protonic zeolites H-Y, H-ZSM-5 and H-FER display the expected correlation between Δν(OH) and ΔH0 for both CO and N2; that is, increasing (absolute) values of Δν(OH) correspond to increasing (absolute) values of ΔH0. However, MWW type zeolites clearly break that trend; they all show a distinctively smaller value of ΔH0 than (for instance) H-FER and H-ZSM-5, and yet a larger value of Δν(OH). Regarding H-MCM-22 (Si:Al = 24.5:1), the smaller ΔH0 value, as compared to H-FER, for the formation of hydrogen-bonded OH⋯CO complexes with the zeolite Brønsted-acid sites was also confirmed by the calorimetric results reported in Section 3.2. Moreover, Li et al.78 and Tsai et al.79 found, by means of temperature programmed desorption of adsorbed ammonia, that (for nearly the same Si:Al ratio) H-MCM-22 has a weaker Brønsted acidity than H-ZSM-5. This is in agreement with the correspondingly smaller values of ΔH0(CO) and ΔH0(N2) shown by H-MCM-22 (Table 3), but not so with its distinctively larger values of Δν(OH)(CO) and Δν(OH)(N2) when compared to H-ZSM-5. It should be clear, therefore, that the rather common practice of ranking Brønsted-acid strength of protonic zeolites according to the corresponding O–H frequency shift probed by an adsorbed weak base can be misleading, particularly when dealing with zeolites not having the same structure type. Nevertheless, within the same structural group Δν(OH) seems to correlate with ΔH0, as shown by H-MCM-22 and H-MCM-56; both of them belonging to the same (MWW) structural group. Note also that within this group of zeolites the Brønsted acid strength, as tested by the corresponding ΔH0(CO) or ΔH0(N2) value, slightly decreases with decreasing Si:Al ratio along the series H-MCM-22 (Si:Al = 24.5:1), H-MCM-22 (Si:Al = 16.4:1), H-MCM-56 (Si:Al = 16:1), as could in principle be expected.
Attempts to calculate the interaction energy between H-MCM-22 and adsorbed CO by DFT calculations did not lead to quantitatively accurate results. Accurate calculations at the coupled cluster level of theory performed for cluster models show that dispersion-corrected DFT methods overestimate the interaction energy by about 50%. In contrast, the standard GGA functional (PBE) gives interaction energies in significantly better agreement with the coupled cluster method (for the cluster model) and with experiment (periodic model). The relatively good performance of the PBE functional is however due to fortuitous error cancellation. O–H and C–O stretching vibrations can be accurately described with the DFT method, only due to the fact that the ω/r correlation scheme used here corrects the DFT results to the coupled cluster accuracy. A highly accurate description of CO adsorption on Brønsted-acid sites remains a challenge for computational chemistry.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/c3cp54738h |
This journal is © the Owner Societies 2014 |