Introduction

In the past few decades, M(IV)-incorporated zeolites have gained enough attention1,2,3. The discovery of Titanium silicate-1 (TS-1) in 19834 has been regarded as a milestone in heterogeneous catalysis5 and other Ti(IV)-containing zeolites such as Ti-MWW, SSZ-24 and Ti-UTL were subsequently developed that also exhibit satisfying catalytic effects for a wide spectrum of substrates6,7,8,9,10. Corma, Přech and their collaborators11,12 reported that Sn-BEA zeolite is an efficient catalyst for the Baeyer-Villiger oxidation reactions. As a matter of fact, Sn- and Zr-BEA zeolites are also known to exhibit superior catalytic performances for the transformation of cellulosic biomass to downstream chemicals and fuels3,13,14,15. Albeit less used for catalytic applications, Ge(IV) ions introduced into the framework of zeolites have the directing effect that results in the formation of ITQ-24, ICP-2 and other new structures, and the ADOR (assembly-disassembly-organization-reassembly) mechanism was posed for the synthesis processes16,17.

Isolated and well-defined M(IV) sites have been acknowledged as the active sites of M(IV)-incorporated zeolites1,2,3,5,13,14,15,18,19,20. As testified by density functional and two-layer ONIOM calculations, M(IV) ions exhibit outstanding adsorption performances for a variety of probe molecules such as H2O, NH3 and amino acids21,22,23,24,25,26,27,28,29,30,31,32,33,34,35,36,37,38,39,40. The Ti-peroxo species in TS-1 zeolite produced from the interaction of Ti(IV) sites and H2O2 are the active sites for alkene epoxidation41,42,43,44,45,46,47,48,49,50,51,52,53,54,55,56,57, and Wells et al.48 demonstrated that Ti(IV) sites adjacent to Si vacancies [(OSiO3)3TiOH] are more reactive; that is, the epoxidation processes are significantly accelerated by defects. The two Ti(IV) species [i.e., Ti(OSiO3)4 and (OSiO3)3TiOH] were identified in TS-1 zeolite49,50 and Sn(IV) analogues were proposed for Sn-BEA zeolite51,52,53. Defects in Sn-BEA zeolite were reported to play a significant promoting effect during the transformation of cellulosic biomass54,55,56,57,58,59,60.

Active sites are one of the central topics for adsorption and catalysis, while a number of key issues remain enigmatic for M(IV)-incorporated zeolites; e.g., we are not clear which causes the differences of active sites in Sn-BEA and TS-1 zeolites, topology of zeolites or identity of M(IV) ions. The active sites of Sn-BEA zeolite were assigned to be octahedral Sn(IV) species composed by tetrahedral Sn(IV) sites and two water molecules61, while the Ti(IV) active sites in TS-1 zeolite prefer to being five-coordinated upon water adsorption62. In this work, periodic density functional theory (p-DFT) calculations were conducted to have a comprehensive understanding of M(IV) active sites and their coordination numbers in M(IV)-incorporated zeolites, considering the effects such as topology of zeolites (BEA, FER, CHA), identity of M(IV) ions (M = Sn, Ti, Zr, Ge), type of framework species (M P , M L , M B , see Fig. 1) and choice of crystallographically distinct T sites: (1) M2(IV) active sites in BEA zeolites containing the various M(IV) ions (M = Sn, Ti, Zr, Ge). The second water that forms only H-bonds promotes chemisorption of the first water and has comparable or even larger adsorption energies; (2) M9(IV) and M6(IV) active sites in BEA zeolites, where Sn(IV) and Zr(IV) active sites are hexa-coordinated while Ti(IV) and Ge(IV) active sites remain penta-coordinated; (3) M(IV)-BEA zeolites with adsorption of three and four water molecules. The third water at M2(IV) sites continues to promote chemisorption of the first water while subsequent adsorption may play an adverse effect. Sn(IV) and Zr(IV) active sites are potentially expanded to hepta-coordination; (4) M(IV)-CHA and M(IV)-FER zeolites. Different from other zeolites, Ti(IV) active sites in FER zeolites prefer to being hexa-coordinated as Sn(IV) active sites; (5) Lewis acidic defects in M(IV)-BEA zeolites, which produce significantly beneficial effects for water adsorption. The coordination number of Ti(IV) sites is expanded to six while the hepta-coordinated Sn(IV) species is inhibited; (6) Brϕnsted acidic defects, which promote chemisorption for the first water while prevents the second water from chemisorption. A second form of Brϕnsted acidic defects with higher stability was detected in Sn-BEA zeolite that shows distinct adsorption properties. Results obtained thus far are beneficial to understand the structural, adsorption and coordination aspects of M(IV)-incorporated zeolites and to decipher the active sites that are critical to adsorption and catalytic processes.

Figure 1
figure 1

Different framework M(IV) species within zeolites that are referred to as (A) perfectly tetrahedral M(IV) sites (M P ), (B) defects with Lewis acidity (M L ) and (C) defects with Brϕnsted acidity (M B ).

Computational Details

Models

Atomic coordinates of zeolites were downloaded from the International Zeolite Association (IZA) website63. The periodic models of M(IV)-BEA, M(IV)-FER and M(IV)-CHA zeolites as used previously39 were displayed in Fig. 2, wherein M(IV)-FER and M(IV)-CHA consist of two and four unit cells along the c (1 × 1 × 2) and a × b (1 × 2 × 2) lattice vectors, respectively. Different from CHA and FER zeolites where all T sites are indistinguishable (referred to as T1), BEA possesses nine crystallographically distinct T sites, and as recommended elsewhere28,29,30,33,35,38,39,53,54,55,56,57,58,59,60,61, T2 site that is the most energetically favorable and T9 and T6 sites that are situated at the intersection of two channels were investigated.

Figure 2
figure 2

Periodic models of (A) BEA, (B) CHA and (C) FER zeolites labeling the T sites presently investigated.

Lewis and Brϕnsted acidic defects in M(IV)-incorporated zeolites were given in Figs 1 and 3. Lewis acidic defects (M L ) were constructed by removing a neighboring Si atom as well as its first-shell three Si atoms26,34,48,49,50,55, while Brϕnsted acidic defects (M B ) were created with the formation of the ≡M(OH)2Si≡ linkage31,53,58,64.

Figure 3
figure 3

Local structures for defects within M(IV)-BEA zeolites that display (A) Lewis acidity (M L ) and (B) Brϕnsted acidity (M B ).

Methods

The Perdew, Burke and Ernzrhof (PBE) exchange-correlation functional65,66 supplemented with the damped C6 dispersion term67 (referred to as DFT-D2, implemented in the VASP software) was used. The standard PAW (projected augmented wave) psuedopotentials were used for all elements, while M(IV) ions are exceptions that were handled by the highest electronic PAW pseudopotentials (Sn_d for Sn, Ti_pv for Ti, Zr_sv for Zr and Ge_d for Ge), in that their semi-core s, p or d states should be regarded as valence electrons33,37,39,55,68. The energy cutoff was 400.0 eV, and the Brillouin zone sampling was restricted to Γ-point. Structural optimizations finished when the forces on each atom are below 0.05 eV Å−1. The adsorption energies of the Nth water within M(IV)-incorporated zeolites were defined as,

$${E}_{{\rm{adN}}}=E({\rm{MZeo}}-{{\rm{nH}}}_{2}{\rm{O}})\mbox{--}E({\rm{MZeo}}\,-\,({\rm{n}}-1){{\rm{H}}}_{2}{\rm{O}})-{E}({{\rm{H}}}_{2}{\rm{O}})$$
(1)

where MZeo-nH2O stands for M(IV)-incorporated zeolites respectively adsorbed with N-numbered water (n = 0, 1, 2, 3, 4). Noting that the E ad2 calculations were based on the lower-energy MZeo-H2O configurations (e.g., Sn9 P b rather than Sn9 P a due to the larger E ad1 value and higher stability, see Table 1).

Table 1 M-OW distances (Å) and adsorption energies of the nth water (E adn, kJ/mol) for defect-free M(IV)-BEA zeolites (M = Sn, Ti, Zr, Ge; n = 1, 2).

Results and Discussion

As illustrated in Fig. 4, water can approach some T sites of zeolites via different directions that are represented by “a”, “b” and “c”. Nomenclature of adsorption configurations includes such information as M(IV) species (M P , M L , M B , see Fig. 1), number of T sites (19) and direction of water adsorption (a, b, c); e.g., M9 P ab (Fig. 5) in BEA zeolite stands for the adsorption configuration where two water molecules approach the perfectly tetrahedral M9(IV) site via “a” and “b” directions (M = Sn, Ti, Zr, Ge). Adsorption configurations where water is assumed to form H-bonds with other water molecules and framework-O atoms (referred to as OF) are suffixed by hi (i = 1, 2, …), see M2 P h1 in Fig. 5 for instance.

Figure 4
figure 4

(A) Two and (B) potential directions present for water molecules to approach the M(IV) sites within zeolites. Different directions are referred to as “a”, “b” and “c” that are colored in semitransparent pink, yellow and green, respectively.

Figure 5
figure 5

Local configurations of water adsorption at defect-free M2(IV), M9(IV) and M6(IV) sites of BEA zeolites, where M = Sn, Ti, Zr, Ge. OW2 in M2 P h1 refers to the water molecule that is stabilized mainly by H-bonds.

M(IV)2 Sites in Defect-free BEA Zeolite

The local adsorption configurations of water at defect-free M2(IV)-BEA zeolites are shown in Fig. 5 (M2 P a), and the M-OWa distances and adsorption energies (E ad1) are given in Table 1. The M-OWa distances are 2.356, 2.365, 2.542 and 2.425 Å respectively for M = Ti, Sn, Ge and Zr, in line with previous reports21,22,23,25,26,27,28,29,30,31,32,33,34,35,36,37,38 and indicative of chemisorption. Although with a smaller radius than Sn(IV) (0.53 Å vs. 0.71 Å), Ge(IV) corresponds to larger M-OWa distances suggesting inferior interactions. The adsorption energies of water (E ad1) are calculated at −44.9, −64.4, −27.9 and −64.8 kJ/mol respectively for M = Ti, Sn, Ge and Zr, and the interaction strengths within Ge2(IV)-BEA zeolite are almost at the level of all-siliceous zeolite (−20.3 kJ/mol). The adsorption performances of M(IV)-BEA zeolites are in line with their catalytic effects (Zr ≈ Sn > Ti > Ge)2,3,13,14,15, including the results of two-dimensional zeolites available69. The thermodynamic changes for the hydrolysis of M-O bonds fall within the range of 63.873.9 kJ/mol, which are obviously lower than that of all-siliceous zeolite (106.2 kJ/mol) and indicate the less stability and more likelihood of forming defects due to M(IV) incorporation.

M(IV)2 sites are situated at the straight channel of BEA zeolite and it is assumed that only one water can be chemisorbed, as testified by adsorption of the second water that forms H-bonds with the first water and OF atoms, see M2 P h1 in Fig. 5. Accordingly, all M(IV)2 active sites are penta-coordinated with adsorption of one water. Strong H-bonds are constructed between two adsorbed water molecules and the OWaH•••OW2 distances are 1.684, 1.607, 1.664 and 1.658 Å respectively for M = Ti, Sn, Ge and Zr. The second water is further stabilized by three H-bonds with OF atoms; e.g., the OW2H•••OF H-bonds in Sn2 P h1 are 2.005, 2.327 and 2.765 Å. Interestingly, its presence promotes the interaction of M2(IV) sites with the first water, since the M-OWa distances are respectively 2.255, 2.267, 2.212 and 2.352 Å for M = Ti, Sn, Ge and Zr and all are shorter than those of M2 P a (Table 1). As a synergy of the two effects, the adsorption energies of the second water in M2 P h1 (E ad2) are substantial and amount to −58.6, −67.1, −53.0 and −62.1 kJ/mol respectively for M = Ti, Sn, Ge and Zr, which are comparable to, if not larger than, the E ad1 values of M2 P a where water is chemisorbed; in addition, the E ad2 values are close for all M(IV) ions and are not so dependent on the identify of M(IV) ions as those of chemisorbed water (E ad1).

M(IV)9 and M(IV)6 Sites in Defect-free BEA Zeolite

M(IV)9 sites in BEA zeolite are at the intersection of two channels and can be approached by water via either “a” or “b” direction, see Figs 4 and 5 (M9 P a and M9 P b)33. For a specific M(IV) ion, the M-OWa and M-OWb distances are close to each other suggesting comparable interactions from two directions, as corroborated by the calculated adsorption energies (E ad1, Table 1). When two water molecules via “a” and “b” directions approach the M9(IV) sites of M9(IV)-BEA zeolites at the same time, both of them construct direct bonds with Sn9(IV) and Zr9(IV) sites while only one remains chemisorbed for Ti9(IV) and Ge9(IV) sites, see Fig. 5 (Sn9 P ab and Ti9 P ab). That is, the coordination numbers of Sn9(IV), Zr9(IV) and Ti9(IV), Ge9(IV) active sites are expanded respectively to six and five, in line with the results of dynamic nuclear polarization surface enhanced NMR spectra of Sn-BEA zeolite and (resonant) valence-to-core X-ray emission spectra of TS-1 zeolite61,62. The adsorption energies of the second water (E ad2) in M9 P ab are calculated to be −19.0, −42.6, −20.5 and −61.0 kJ/mol respectively for M = Ti, Sn, Ge and Zr, consistent with the changing trends of M-OW distances (Table 1). Owing to competition of two chemisorbed water, the E ad2 values in Sn9 P ab and Zr9 P ab reduce to some extent as compared to the E ad1 values, while the E ad2 value of Ti9 P ab due to mainly H-bonding interactions descends substantially and is close to those of all-siliceous zeolite (Table 1).

M(IV)6 sites at the intersection of BEA zeolite are also investigated that substantialize the results of M9(IV) sites: Sn6(IV) and Ti6(IV) sites can chemisorb at most two and one water molecules, respectively, see their local structures in Fig. 5 (Sn6 P ab and Ti6 P ab) and geometric and energetic data in Table 1. That is, Sn6(IV) active sites adopt the octahedral geometry with adsorption of two water molecules while Ti6(IV) active sites are penta-coordinated with adsorption of one water.

Adsorption with More Water Molecules

Figure 6 depicts the local configurations of Sn2(IV)-BEA zeolite adsorbed with three (Sn2 P h2) and four (Sn2 P h3) water molecules. In Sn2 P h2, the second and third water molecules are located at either side of the chemisorbed one and their chemical environments resemble each other. The OWaH•••OW3 and three OW3H•••OF H-bonds are 1.603 Å and 2.134, 2.163, 2.961 Å, respectively. The Sn2-Owa distance in Sn2 P h2 equals 2.186 Å and is further shortened than in Sn2 P h1 (2.267 Å); nonetheless, E ad3 (−64.2 kJ/mol) decreases slightly as compared to E ad2, due to the inferior H-bonding interactions as reflected by their distances. The E ad4 value (−33.5 kJ/mol) of Sn2 P h3 shows considerable reductions albeit the fourth water also constructs H-bonds with the chemisorbed water, due to the obviously weaker H-bonding interactions; e.g., the OW4H•••OWa distance equals 1.971 Å and is apparently longer than those of the second and third water molecules. Subsequent adsorption constructs the second, third and higher-order water shells around the chemisorbed one, which will not be discussed here.

Figure 6
figure 6

Local adsorption configurations for three and four water molecules within defect-free Sn2(IV)-BEA zeolite. OW3 and OW4 refer to the water molecules that are stabilized mainly by H-bonds.

H-bonded adsorption configurations of Sn9(IV) site are shown in Fig. 7 (Sn9 P h1 and Sn9 P h2) that resemble the condition of Sn2(IV) site. The third water in Sn9 h1 rather than in Sn9 h2 is more stabilized by H-bonds, and the OWaH•••OW3 and OWbH•••OW3 distances are optimized at 1.677 and 2.151 Å, respectively. The third water in Sn9 h1 facilitates the chemisorption of associated water (Sn9-OWa: 2.289 Å in Sn9 h1 vs. 2.402 Å in Sn9 P ab) while in Sn9 P h2 plays an adverse effect for the chemisorption of associated water (Sn9-OWb: 2.425 Å in Sn9 P h2 vs. 2.354 Å in Sn9 P ab). As a result, the adsorption energies of the third water (E ad3) differ markedly for Sn9 P h1 and Sn9 P h2 that amount to −61.2 and −26.9 kJ/mol, respectively. It is interesting to find that three water molecules can also be simultaneously chemisorbed at Sn9(IV) site with elongated Sn-OW distances (2.665, 2.558 and 2.409 Å)70, see Sn9 P abi in Fig. 7 and Table 2. However, the E ad3 value of Sn9 P abi equals −26.6 kJ/mol and is apparently less than that of Sn9 h1. That is, the coordination number of Sn9(IV) site in BEA zeolite should predominate as six although potentially expanded to seven. Parallel p-DFT calculations are conducted for Zr9(IV)-BEA zeolite and it indicates that its coordination number can also be expanded to seven with the Zr-Ow distances of 2.518, 2.518 and 2.454 Å (Table 2). The E ad3 difference of Zr9 P abi vs. Zr9 P h1 (−36.5 vs. −59.8 kJ/mol) is less than that of Sn9 P abi vs. Sn9 P h1, suggesting that Zr9(IV)-BEA zeolite has a greater possibility of developing the hepta-coordination mode.

Figure 7
figure 7

Local adsorption configurations for three water molecules within defect-free Sn9(IV)-BEA zeolite. OWi in Sn9 P abi refers to the water molecule in-between OWa and OWb, and OW3 in Sn9 P h1 and Sn9 P h2 stands for the water molecule that is stabilized mainly by H-bonds.

Table 2 M-OW distances (Å), adsorption energies of the 3th water (E adn, kJ/mol) for M9(IV)-BEA zeolites (M = Sn, Zr).

Defect-free CHA and FER Zeolites

Figure 8 and Table 3 indicate that water can approach M(IV) sites of CHA zeolite from three different directions (“a”, “b” and “c” as shown in Fig. 4). Each direction in Sn(IV)-CHA zeolite results in the formation of direct Sn-OW bonds (Sn-OWa: 2.492 Å; Sn-OWb: 2.395 Å; Sn-OWc: 2.443 Å) while only two directions in Ti(IV)-CHA zeolite can be chemisorbed with water (Ti-OWa: 2.905 Å; Ti-OWb: 2.371 Å; Ti-OWc: 2.414 Å)21,22,23,25,26,27,28,31,32,33, where the Ti-OWa distance is apparently longer. The adsorption energies of three directions (E ad1) differ remarkably and are maximal for M1 P b with the values being −72.3 and −57.1 kJ/mol for M = Sn and Ti, respectively (Table 3). For adsorption of two water molecules in Sn(IV)-CHA zeolite, Sn1 P ab remains hexa-coordinated (Sn-Ow: 2.457 and 2.261 Å) while Sn1 P ac (Sn-Ow: 2.565 and 3.033 Å) and Sn1 P bc (Sn-Ow: 2.307 and 3.524 Å) transform to penta-coordination. Instead, all adsorption configurations of Ti(IV)-CHA zeolite are presented as penta-coordination including the one corresponding to Sn1 P ab (i.e., Ti1 P ab with Ti-Ow distances of 3.699 and 2.347 Å). Chemisorption of three water molecules in Sn(IV)-CHA zeolite seems impossible, see the optimized structure (Sn1 P abc) in Fig. 8. In consequence, Sn(IV) and Ti(IV) active sites of CHA zeolites are respectively hexa- and penta-coordinated that resemble the condition of BEA zeolites.

Figure 8
figure 8

Local configurations for water adsorption within defect-free M(IV)-CHA zeolites, where M = Sn, Ti.

Table 3 M-OW distances (Å), adsorption energies of the nth water (E adn, kJ/mol) for defect-free M(IV)-CHA and M(IV)-FER zeolites (M = Sn, Ti; n = 1, 2).

Figure 9 shows the adsorption configurations of water within M(IV)-FER zeolites. Water can be chemisorbed at the M(IV) sites of FER zeolite from two different directions referred to as “a” and “b”, and the Sn-Owa, Ti-Owa, Sn-Owb and Ti-Owb distances in Sn1 P a, Ti1 P a, Sn1 P b and Ti1 P b are respectively 2.398, 2.392, 2.398 and 2.449 Å. Chemisorption of two water molecules is viable in Sn(IV)-FER zeolite as in Sn(IV)-BEA and Sn(IV)-CHA zeolites (Sn-Ow: 2.379 and 2.392 Å), implying that chemisorption of the first water facilitates the interaction of the second water as verified subsequently. It is surprising to find that two water molecules can be chemisorbed at the Ti(IV) sites of FER zeolite, and the Ti-Owa and Ti-Owb distances in Ti1 P ab are 2.365 and 2.419 Å that are also shorter than those adsorbed with one water (Table 3). Accordingly, both Sn(IV) and Ti(IV) active sites in FER zeolites adopt the octahedral geometry with adsorption of two water molecules. As indicated in Table 3, the adsorption energies of the second water (E ad2) are comparable to those of the first water (E ad1) from the same direction (e.g., “b” direction for Ti1 P ab and Ti1 P b).

Figure 9
figure 9

Local configurations for water adsorption within defect-free M(IV)-FER zeolites, where M = Sn, Ti.

Figure 4A shows that “a” and “b” directions each have three different OFOFOFM dihedrals, and the accessibility of water to M(IV) sites is estimated by the minimal dihedral33,

$${{\rm{\Omega }}}_{{\rm{\min }}}=\,{\rm{\min }}\,[{\rm{\Omega }}({{\rm{O}}}_{{\rm{F}}1}{{\rm{O}}}_{{\rm{F}}2}{{\rm{O}}}_{{\rm{F}}3}{\rm{M}}),{\rm{\Omega }}({{\rm{O}}}_{{\rm{F}}2}{{\rm{O}}}_{{\rm{F}}3}{{\rm{O}}}_{{\rm{F}}1}{\rm{M}}),{\rm{\Omega }}({{\rm{O}}}_{F3}{{\rm{O}}}_{{\rm{F}}1}{{\rm{O}}}_{{\rm{F}}2}{\rm{M}})]$$
(2)
$${{\rm{\Psi }}}_{{\rm{\min }}}=\,{\rm{\min }}\,[{\rm{\Psi }}({{\rm{O}}}_{{\rm{F}}1}{{\rm{O}}}_{{\rm{F}}4}{{\rm{O}}}_{{\rm{F}}2}{\rm{M}}),{\rm{\Psi }}({{\rm{O}}}_{{\rm{F}}4}{{\rm{O}}}_{{\rm{F}}2}{{\rm{O}}}_{{\rm{F}}1}{\rm{M}}),{\rm{\Psi }}({{\rm{O}}}_{{\rm{F}}2}{{\rm{O}}}_{{\rm{F}}1}{{\rm{O}}}_{{\rm{F}}4}{\rm{M}})]$$
(3)

Chemisorption of one water transforms the M(IV) geometry from tetrahedral to bipiramidal21,23,33, and the dihedral of direction with water adsorption reduces substantially while that of the other direction ascends considerably. Chemisorption of the second water causes the Ωmin and Ψmin values to again get close to each other and reduce pronouncedly due to the formation of octahedral geometry61, see the Ωmin and Ψmin values for Sn9(IV) sites of BEA zeolite in Table 4. In M(IV)-FER zeolites, however, both Ωmin and Ψmin decline due to adsorption of the first water although to an apparently less extent for the other direction with no water (Table 4), and the smaller dihedral of the other direction indicates that adsorption of the first water facilitates the accessibility of the second water towards M(IV) sites, which is distinct from the condition of other zeolites. It thus deciphers why two water molecules can be chemisorbed at Ti(IV) sites of FER rather than other zeolites.

Table 4 The OFOFOFM dihedrals corresponding to the “a” (Ωmin, degrees) and “b” (Ωmin, degrees) directions for defect-free M(IV)-incorporated zeolites adsorbed with water molecules (M = Sn, Ti; n = 0, 1, 2)a.

Lewis Acidic Defects

Defects have been implicated to be critical for catalytic reactions54,55,56,57,58,59,60, and is this effect associated, at least in part, with alteration of the active sites? The adsorption configurations of water at M9(IV) Lewis acidic defects (M9 L ) of BEA zeolites are shown in Fig. 10. Water can approach M9(IV) Lewis acidic defects from “a” and “b” directions as in the condition of perfectly tetrahedral sites (M9 P ) while interactions are reinforced (Table 5); e.g., the adsorption energies (E ad1) are −85.1 and −89.1 kJ/mol respectively for Sn9 L a and Sn9 L b and surpass those of Sn9 P a and Sn9 P b (Table 1)49,55. More significant promotion effects of Lewis acidic defects are detected during adsorption of the second water: The E ad2 value in Sn9 L ab equals −78.6 kJ/mol and is even larger than the E ad1 values of perfectly tetrahedral sites. Two water molecules can be chemisorbed at Ti9(IV) Lewis acidic defects (Ti-OW distances of Ti9 L ab: 2.278 and 2.233 Å), and the E ad2 value (−55.0 kJ/mol) surpasses the E ad1 values of perfectly tetrahedral Ti9(IV) sites (Table 1), which agree with the results of Sn9(IV) Lewis acidic defects. The adsorption configurations of three water molecules at Sn9(IV) Lewis acidic defects (Fig. 11) are close to those of perfectly tetrahedral sites, and the E ad3 values are calculated at −16.6 and −66.0 kJ/mol for Sn9 L abi and Sn9 L h1, respectively (Table 2). Accordingly, Lewis acidic defects inhibit somewhat the formation of hepta-coordinated Sn9(IV) species (Chemisorption of three water molecules is also tried for Ti9(IV) Lewis acidic defects, while only two water molecules remain bonded and the third water forms H-bonds with other water molecules and OF atoms, see Fig. 11 (Ti9Labi)).

Figure 10
figure 10

Local adsorption configurations for one and two water molecules at M9(IV) Lewis acidic defects (M9 L ) of BEA zeolites, where M = Sn, Ti.

Table 5 M-OW distances (Å) and adsorption energies of the nth water (E adn, kJ/mol) for M9(IV) Lewis acidic defects of BEA zeolites (M = Sn, Ti; n = 1, 2).
Figure 11
figure 11

Local adsorption configurations for three water molecules M9(IV) Lewis acidic defects (M9 L ) of BEA zeolites, where M = Sn, Ti. OWi in Sn9 L abi refers to the water molecule in-between OWa and OWb, and OW3 in Sn9 L h1 and Sn9 L h2 stands for the water molecule that is stabilized mainly by H-bonds. In Ti9 L abi, three water molecules are assumed to form direct bonds as in Sn9 L abi.

Table 6 lists the root-mean-square deviations (RMSD) of local M(IV) sites during water adsorption. For each M(IV) site, RMSDs generally increase with the number of chemisorbed water molecules, implying larger structural perturbations; in addition, at a specific chemisorbed water content, Lewis acidic defects always result in obviously higher RMSDs than corresponding perfectly tetrahedral sites; e.g., the RMSDs are calculated to be 0.20 and 0.24 Å in Sn9 P a and Sn9 P ab while are enlarged to 0.34 and 0.47 Å in Sn9 L a and Sn9 L ab, respectively. Accordingly, the more structural flexibility of Lewis acidic defects facilitates the interaction with water and allows the formation of hexa-coordinated Ti(IV) species in BEA zeolite.

Table 6 Root-mean-square deviations (RMSD, Å) for M9(IV) active sites of BEA zeolites due to water adsorption (M = Sn, Ti)a,b.

Brϕnsted Acidic Defects

The adsorption configurations of NH3 at defect M9 B of BEA zeolite with the ≡M(OH)2Si≡ linkage (Figs 1 and 3) are given in Fig. 12. Defect M9 B in BEA zeolite transfers the proton to NH3 automatically forming NH4 + and thus shows Brϕnsted acidity, which are in line with the results of M(IV)-incorporated MFI zeolites32,64. The bridging hydroxyls of defect M B significantly accelerate the isomerization reaction of glucose to fructose58. Figure 13 depicts the adsorption configurations of water at M9(IV) Brϕnsted acidic defects of BEA zeolites, with structural parameters and adsorption energies (E adn) being listed in Table 7. The M-OB and Si-OB distances are altered but not markedly during water adsorption. The Si-OB distances are obviously elongated than in perfectly tetrahedral Si(IV) sites (ca. 1.62 Å) while M(IV) sites remain tightly bonded with two bridging OB atoms. Sn9 B a is an exception where the Si-OB2 bond is ruptured and the Sn-OB distances reduce considerably. The M-OW distances in M9 B b are shorter than in perfectly tetrahedral M(IV) sites and hence interactions with water are enhanced as verified by adsorption energies (e.g., E ad1 = −91.3 kJ/mol for Sn9 B b vs. −74.2 kJ/mol for Sn9 P b). Formation of Brϕnsted acidic defects (M B ) facilitates the accessibility of water to M(IV) sites as reflected by the significantly smaller dihedrals than in perfectly tetrahedral sites: The OF3OB1OB2M dihedrals (Fig. 3) are 10.06° and 10.78° respectively for Sn9 B and Ti9 B . However, limited space is available for the second water and as a result, water from “a” direction forms only H-bonds instead of being chemisorbed: The OB1HB1•••OWa and OWaH•••OF H-bonds are respectively 1.499 and 1.918, 2.836, 2.670 Å in Sn9 B a while 1.662 and 2.139, 2.573 Å in Ti9 B a, and stronger OB1HB1•••OWa H-bond in Sn9 B a results from higher Brϕnsted acidity32 that agrees with larger OB1-HB1 distances (1.053 vs. 1.011 Å). The E ad1 values are −101.3 and −69.2 kJ/mol respectively for Sn9 B a and Ti9 B a, and structural distortions in Sn9 B a may facilitate the interactions as testified subsequently. For both Sn9 B ab and Ti9 B ab, one water is chemisorbed while the other is H-bonded.

Figure 12
figure 12

Local configurations for NH3 adsorption at two types of Sn9(IV) Brϕnsted acidic defects (Sn9 B and Sn9 B ) and of Sn-BEA zeolite.

Figure 13
figure 13

Local configurations for water adsorption at M9(IV) Brϕnsted acidic defects (M9 B ) of BEA zeolites, where M = Sn, Ti. There emerges a second Sn(IV) Brϕnsted acidic defect named Sn9 B , where OB2 forms direct bonds only with the Sn(IV) sites. Both Sn9 B ab and Sn9 B ba are adsorption configurations with two water molecules while the chemisorbed water are from the “a” and “b” directions, respectively.

Table 7 M-OW, M-OB and Si-OB distances (Å) and adsorption energies of the nth water (E adn, kJ/mol) for M9(IV) Brϕnsted acidic defects (M9 B) of BEA zeolites (M = Sn, Ti; n = 1, 2)a.

The Si-OB2 bond in Sn9 B a is broken, and such defect referred to as Sn9 B also exists in Sn(IV)-BEA zeolite. In addition, Sn9 B is more stable than Sn9 B and their energy difference equals −22.8 kJ/mol; however, Ti9 B is non-existent and spontaneously transforms to Ti9 B . The OB2HB2 group in Sn9 B is flexible and allows water chemisorption from either “a” or “b” direction, see Sn9 B a and Sn9 B b in Fig. 13. Chemisorption of water from “b” direction causes structural reconstruction to resemble Sn9 B b. The adsorption energies (E ad1) in Sn9 B a and Sn9 B b are respectively −8.5 and −68.5 kJ/mol, and the particularly small value in Sn9 B a is caused by the serious structural distortion in order to accommodate the chemisorbed water. Accordingly, water from “a” direction of Sn9 B should be preferentially the H-bonded adsorption configuration (i.e., Sn9 B a), and the E ad1 value calculated this way amounts to −78.5 kJ/mol (Table 7). Similar to the condition of Sn9 B , only one water can be chemisorbed at Sn9 B , see Fig. 13 (Sn9 B ab and Sn9 B ba where water is chemisorbed from “a” “b” directions, respectively). Accordingly, M(IV) Brϕnsted acidic defects, for both acidic forms and different M(IV) ions, result in only the hexa-coordination mode with chemisorption of one water.

Conclusions

Periodic density functional theory (p-DFT) calculations have been used to comprehensively investigate the active sites of M(IV)-incorporated zeolites, considering the identity of M(IV) ions, topology of zeolites, type of framework species and choice of T sites.

With regard to defect-free BEA zeolites, all M(IV) active sites are penta-coordinated with chemisorption of one water when situated at the straight channel while divergences arise when situated at the intersection: Sn(IV) and Zr(IV) active sites predominate as hexa-coordination while Ti(IV) and Ge(IV) active sites remain penta-coordinated; in addition, it is surprising to find that Sn(IV) and Zr(IV) are potentially expanded to hepta-coordination although with relatively small probabilities. For M2(IV) sites, the second and third water molecules form strong H-bonds with chemisorbed water and framework-O atoms and promote the interaction of chemisorbed water. The adsorption energies of the second water, irrelevant of the identity of M(IV) ions, are comparable to, if not larger than, those of chemisorbed water, while those of the third and fourth water molecules, especially the latter, show reduction.

Results of CHA zeolites, where water can approach M(IV) sites from three directions, are similar to those of BEA zeolites: Sn(IV) active sites are hexa-coordinated while Ti(IV) active sites are penta-coordinated. Sn(IV) active sites in FER zeolite adopt the hexa-coordination mode as in the condition of other zeolites, while it is surprising to find that Ti(IV) active sites are also presented as hexa-coordination. Chemisorption of the first water at M(IV) sites of FER zeolites facilitates the interaction with the second water, as verified by the dihedral analyses.

Owing to enhanced structural flexibility, Lewis acidic defects reinforce the adsorption of water and the promoting effects are more obvious during chemisorption of the second water; in addition, Ti(IV) Lewis acidic defects can be expanded to hexa-coordination, while hepta-coordinated Sn9(IV) species is somewhat inhibited. M(IV) Brϕnsted acidic defects also facilitate the interaction of the first water while present the second water from chemisorption. A second form of Brϕnsted acidic defects that has higher stability exists in Sn(IV)- rather than Ti(IV)-BEA zeolites. Two forms of Sn(IV) Brϕnsted acidic defects show divergent adsorption properties and can be inter-converted during water adsorption. Despite that, all M(IV) Brϕnsted acidic defects are hexa-coordinated, irrespective of different M(IV) ions or acidic forms, due to limited space available for the second water.