Density functional study on the bimetallic TimZrn (n + m ≤ 5) clusters and their interactions with H2*

Project supported by the Scientific Research Plan Foundation of Sichuan Education Department of China (Grant No. 2014JY0072).

Zhang Ge, Sheng Yong
College of Materials Science and Engineering, Sichuan University, Chengdu 610065, China

 

† Corresponding author. E-mail: shengyong69@163.com

Project supported by the Scientific Research Plan Foundation of Sichuan Education Department of China (Grant No. 2014JY0072).

Abstract

Equilibrium geometries, stabilities, and electronic properties of small TimZrn (n + m ≤ 5) clusters were investigated using the density functional method. The ground states were determined, and it was found that the larger clusters and those consisting of more Zr atoms are more stable. The electronic properties of the clusters were discussed based on HOMO–LUMO gaps, vertical ionization potentials (VIP), and vertical electron affinities (VEA). Furthermore, we studied the interactions between those clusters and molecular hydrogen, and found that in all the cases dissociative chemisorptions occurred. According to the chemisorption energies, the pure Zr clusters are relatively more active towards H2 when compared with the others except Ti3Zr, which shows the highest activity. The magnetic moments of TimZrn and TimZrnH2 were also compared, and the results show that the hydrogenated clusters have the same or decreased total magnetic moments with respect to the bare clusters except for Ti3Zr2.

1. Introduction

In the past few decades, a great deal of attention has been paid to metallic clusters due to their unique physical and chemical properties, which can be drastically different from those of their bulk counterparts.[14] Investigations on clusters can also build a bridge for people to understand the evolution from atoms to bulk systems. With vast application prospects, studying the atom aggregation process and cluster stability is of great significance, about which theoretical methods can provide more details compared with experimental approaches.[57] While studying pure atomic clusters leads to fundamental understanding of the quantum size effects on the properties of the clusters, bimetallic clusters consisting of two different atoms have been found to have properties that differ from those of the pure constituent atom clusters.[812] Furthermore, by changing the cluster size and the atom number, some properties of a cluster may be adjusted for different applications.

As transition metal (TM) elements, zirconium and titanium have been extensively explored as structural and functional materials. Alloys consisting of Ti and Zr have huge potential in many fields such as biomedicine,[13] dentistry,[14] catalysts,[15] and storage systems.[16,17] Titanium and zirconium are 3d and 4d TM elements, which belong to the same family, with the outer electron configurations of 3d24s2 and 4s24p64d2, respectively. Closeness and spatial overlap of d orbitals along with s and p orbitals usually make their interatomic reactions complicated.[18] Concerning their nanoclusters, many studies have been focused on them under the framework of the density functional theory (DFT). The evolutions of monoatomic Ti and Zr clusters have been extensively studied,[1924] and their stable geometries, energetic, and electronic properties were the main topics. Those reports have reached the consensus that Ti7, Ti13, Ti15, and Zr7 are magic clusters. Moreover, it was found that many small clusters possess net spin magnetic moments, which is thought to be a common phenomenon for TM clusters. As a further step, some researchers have researched the doped and mixed Ti and Zr clusters using first-principle theories.[2529] Besides, the investigations concerning the interactions of hydrogen molecules with the two kinds of bare clusters revealed that H2 prefers a dissociative chemisorption on them.[3033] Many similar theoretical studies on the adsorption or chemisorption of H2 onto small clusters have been reported, which help people understand the catalytic mechanisms on surfaces and contribute to the development of clusters in hydrogen storage materials.[3437]

However, studies of bimetallic Ti–Zr clusters have been rarely reported. To the best of our knowledge, only Wang et al.[38] carried out a fundamental investigation on (TiZr)n (n = 1–7) clusters, and obtained their stable structures, bonding properties, infrared spectra, and so on. No study on Ti–Zr cluster systems with varied chemical compositions is available. Therefore, in this paper, we present a systematic study of the TimZrn (n + m ≤ 5) clusters. Furthermore, the chemisorption behaviors of H2 on them are studied, which we hope can provide information for hydrogen storage applications.

This paper is organized as follows. In Section 2, we briefly describe the computational methods employed in this work. The results and discussion are presented in Section 3. Finally, we present the main conclusions in Section 4.

2. Computational details

The geometry optimizations and vibrational frequency analyses of bare and hydrogenated TimZrn (n + m ≤ 5) clusters were performed by using the DFT method as implemented in the Gaussian program series. First, a variety of possible initial geometries of every cluster were established in GuassView 3.07. Then, those geometries were optimized in Gaussian 03, and the structures would change to the stable ones. The stable structures were obtained in the absence of imaginary frequency by computing the vibrational frequencies at the same theory level. Besides, all the geometries were optimized at different possible spin states. The lowest-energy structure of a certain cluster was defined as the ground state cluster.

It is of great significance to choose a relatively accurate method to do computational calculations. In the present work, we adopted def2-TZVP,[39] a triple-zeta basis set, to treat all the atoms. To find an appropriate functional, we optimized Ti2, Zr2, and TiH dimers with a variety of exchange–correlation functionals, and compared the results with the values derived from experiments[4043] as listed in Table 1. As we can see, the PBE method can present reasonable results for the concerned indicators, and thus it was chosen to calculate the concerned clusters.

Table 1.

Calculated and experimental equilibrium distances (R), vibrational frequencies (ω), and dissociation energies (ED) of the diatomic clusters.

.
3. Results and discussion
3.1. Structural properties

We first optimized the geometries of the bare TimZrn (n + m ≤ 5) clusters. The most stable structures together with their spin multiplicities and point groups are shown in Fig. 1. Table 2 shows some calculated results obtained from the ground state structures. The TiZr dimer has a bond length of 2.100 Å, which is between the bond lengths of Ti–Ti (1.896 Å) and Zr–Zr (2.271 Å) in the corresponding monoatomic dimers. For the trimers, Ti3 and Zr3 are both equicrural triangles and have the same symmetry of C2v, while Ti2Zr and Zr2Ti are of the Cs symmetry, and the distances between the atoms of the two structures are different. When one Ti atom in the Ti3 cluster is substituted by a Zr atom, the two bonds connecting it are elongated from 2.292 Å to 2.524 Å, and the Ti–Ti bond away from the changed atom is shortened from 2.407 Å to 2.163 Å. By contrast, the Zr2T cluster has two shorter Ti–Zr bonds and a longer Zr–Zr bond after the Ti atom takes the place of the Zr atom in the Zr3 cluster. All the tetramers are tetrahedrons. While the pure clusters are of higher symmetries (C2v for Zr4 and Cs for Ti4), the mixed clusters possess the C1 symmetry. As for the clusters with n + m = 5, the geometries of the most stable structures are all trigonal bipyramids. We also present the structures with energies slightly higher than the minimum-energy ones for comparing. We can see that for the same chemical compositions, the differences of the two most stable structures lie in the atomic arrangements rather than the geometries. Another interesting phenomenon is that to lower the energies of the systems, the Zr atoms tend to occupy the apexes of these trigonal bipyramids, while the Ti atoms are more likely to be located at the middle planes.

Fig. 1. (color online) Equilibrium structures, spin multiplicities, and point group symmetries of the TimZrn (n + m ≤ 5) clusters. The gray balls and blue balls represent Ti and Zr atoms, respectively.

From the above, the TimZrn (n + m ≤ 5) clusters can be seen as deriving from the Zr atoms taking the places of the original Ti atoms in the titanium clusters or vice versa. This may imply that the two metal elements are compatible and substitutable in some cases.

Now we discuss the structural stability, which can usually be described by the binding energy per atom (Eb). The Eb is calculated according to

where E(Ti), E(Zr), and E(TimZrn) are the total energies of the Ti atom, the Zr atom, and the TimZrn cluster, respectively. Figure 2 shows the results. It can be seen that for both pure clusters and mixed clusters, the binding energy per atom increases as the cluster size increases, indicating that these clusters become more stable during the growth process. For the clusters of the same size, it is notable that the binding energy increases monotonically with the increase of Zr atoms, and the changes are almost in a linear fashion. The reason is likely to be that the geometric structures of these same-sized clusters are highly similar, and the outer electron configurations of Ti and Zr are 3d24s2 and 4d25s2, respectively, which are also comparable. The lowest and highest values of Eb appear at the pure Ti clusters and Zr clusters, respectively. Thus, doping Zr atoms in Ti clusters may help to stabilize the clusters.

Fig. 2. (color online) Variations of binding energies as a function of the number of Zr atoms for the TimZrn (n + m ≤ 5) clusters.
Table 2.

Binding energies (Eb), HOMO, LUMO, HOMO–LUMO gaps (Eg), VIP, and VEA of the TimZrn (n + m ≤ 5) clusters in units of eV.

.

The dissociation energy is another indicator to describe the stability of clusters, which is defined as follows:

The ΔE(TimZrn−Ti) and ΔE(TimZrn−Zr) are the energies needed to dissociate a titanium atom and a zirconium atom, respectively, from the TimZrn cluster. Figure 3 presents the variation of ΔE(TimZrn−Ti) with the change of the titanium atom number, and figure 4 presents the variation of ΔE(TimZrn–Zr) with the change of the zirconium atom number. Comparing the two graphs, we can see that for the same cluster, more energy is needed to separate a Zr atom than to separate a Ti atom, which means that Zr atoms take a more important role in the stability of the studied TimZrn clusters. Seeing Figs. 3 and 4 separately, it is found that as the number of Zr atoms increases, i.e., as the number of Ti atoms decreases, both ΔE(TimZrn−Ti) and ΔE(TimZrn−Zr) increase, indicating that the TimZrn cluster gets more stable when the Ti atoms are replaced by Zr atoms. The rule is exactly consistent with that of the averaged binding energies. For the clusters with different sizes, on the whole, the larger clusters (n + m = 4 and 5) present the higher dissociation energies towards both a Ti atom and a Zr atom when compared with the smaller clusters (n + m = 2 and 3), which indicates that the larger TimZrn clusters are characterized with the higher stability. In general, in larger clusters, interactions between atoms are stronger and more complicated, and thus the atoms bind with each other closely. As a result, the clusters are likely to possess higher stabilities.

Fig. 3. (color online) Variations of dissociation energies towards a Ti atom as a function of the number of Ti atoms for the TimZrn (n + m ≤ 5) clusters.
Fig. 4. (color online) Variations of dissociation energies towards a Zr atom as a function of the number of Zr atoms for the TimZrn (n + m ≤ 5) clusters.
3.2. Electronic properties

The energy difference between the highest occupied molecular orbital (HOMO) and the lowest unoccupied molecular orbital (LUMO), i.e., the HOMO–LUMO gap, reflects the ability of an electron to be excited from an occupied orbital to an unoccupied orbital. Generally speaking, a larger HOMO–LUMO gap is related to enhanced chemical stability. The variation of the HOMO–LUMO gaps of the clusters versus the number of zirconium atoms is plotted in Fig. 5. The HOMO–LUMO gaps for the studied clusters range from 0.211 eV to 0.713 eV, indicating that these clusters have a semiconductive feature. The gap changes markedly with the cluster size, and the clusters with n + m = 5 generally possess lower HOMO–LUMO gaps, which we consider results from their relatively low LUMO energies (see Table 2). In regard of the clusters of the same sizes, we can notice that the minima are always found at the pure zirconium cluster, and the clusters consisting of more Ti atoms are more likely to possess large gaps. Besides, for the bimetallic clusters, apart from Ti2, the cluster of Ti2Zr possesses a significantly high HOMO–LUMO gap, indicating that it has better chemical stability.

Fig. 5. (color online) HOMO–LUMO energy gaps as a function of the number of Zr atoms for the TimZrn (n + m ≤ 5) clusters.

The ionization potential and electron affinity are another two significant parameters for the evaluation of chemical stability. In the present work, we have obtained the vertical ionization potentials (VIP) and the vertical electron affinities (VEA) of the TimZrn clusters, which are respectively calculated by

where E(TimZrn) represents the total energy of the neutral cluster, and and denote the total energies of the cationic and anionic species, respectively, with the same geometries of the corresponding neutral clusters. Generally speaking, when the VIP is larger, more energy should be required to remove an electron from the neutral structure without any structural relaxation, and a larger VEA value means that more energy is released when charging the neutral cluster with an extra electron. The VIP and VEA of the considered clusters are presented in Figs. 6 and 7, respectively. The dimers possess the markedly high VIP, while the tetramers are characterized with the lowest VIP. On the whole, when the numbers of atoms are the same, the clusters consisting of more Ti atoms tend to have the higher VIP. As for their behaviors of VEA, the minima can be found at the dimers. The clusters with n + m = 5 possess the largest VEA, implying that they are of lower chemical stability, which is consistent with the fact that their HOMO–LUMO gaps are small. In fact, it is believed that people can predict VEA from LUMO energy, and in this work, it can be verified from the low LUMO energies of these pentamers. The overall trend for each series is opposite to that of VIP, indicating that for the neutral TimZrn clusters, which are less likely to lose an electron (larger VIP), it is also more difficult to charge them with an electron (smaller VEA), thus being chemically stable. Apart from the dimers, one can see that Ti2Zr is less active, which is consistent with the case of the HOMO–LUMO energy gap.

Fig. 6. (color online) Vertical ionization potentials as a function of the number of Zr atoms for the TimZrn (n + m ≤ 5) clusters.
Fig. 7. (color online) Vertical electron affinities as a function of the number of Zr atoms for the TimZrn (n + m ≤ 5) clusters.
3.3. Chemisorption of H2 on the TimZrn (n + m ≤ 5) clusters

One of the main motives of our current work is to investigate the chemical activity of the TimZrn (n + m ≤ 5) clusters towards molecular hydrogen. We performed an exhaustive search in order to find the minimum-energy structures for the hydrogenated clusters. A hydrogen molecule was introduced to each cluster from various directions and distances, and the stable TimZrn structures in both the lowest and the second lowest energies were used in this step. The lowest-energy structures after hydrogenation as well as their spin states and point groups are displayed in Fig. 8. We also present the chemisorption energies, the H–H distances, and the net charges derived from the natural population analysis (NPA) in Table 3. We find that in some cases the ground states of TimZrnH2 are obtained from the metastable structures of the bare TimZrn clusters. The H–H distances vary from 2.182 Å to 3.770 Å, which are far longer than the equilibrium bond length of H2 (0.752 Å). Thus, the molecular hydrogen should be considered as dissociatively chemisorbed onto the considered clusters. Slight structural relaxations of the metallic clusters upon H2 chemisorption were observed. For n + m = 2, 3, 4, the dissociated H atoms prefer bridge sites of the Ti–Zr clusters, except for the Ti4H2 system, in which one H atom is located at the bridge site and the other is on the top of one face. Also, in each dimer-H2 and trimer-H2 system, the atoms are not located in the same plane. When it comes to n + m = 5, the H atoms are on the tops of the faces, except for Ti3Zr2, in which the H atoms are located at the bridge sites. The top sites are not practical in all the studied clusters as the hydrogen molecule is always pushed away from the clusters and there exist imaginary frequencies in the optimized systems. Thus, at least two metallic atoms in each cluster are involved in the interaction with H atoms.

Fig. 8. (color online) Equilibrium structures, spin multiplicities, and point group symmetries of the hydrogenated TimZrn (n + m ≤ 5) clusters. The big gray balls, big blue balls, and small gray balls represent Ti, Zr, and H atoms, respectively.
Table 3.

Chemisorption energies towards H2ECE), H–H distances (RH−H), and charges of the Ti–Zr host clusters and H atoms for the TimZrnH2 (n + m ≤ 5) clusters.

.

The activity of H2 on these clusters can be investigated by calculating the chemisorption energies for the systems according to

where E(H2) and E(TimZrnH2) denote the total energies of H2 and the hydrogenated TimZrn cluster, respectively. The chemisorption energies range from 1.289 eV to 2.947 eV, and the variation is clearly shown in Fig. 9. We notice that the pure Zr clusters are more active towards H2 when compared with the Ti counterparts, characterized by the higher chemisorption energies. Most of the clusters possess chemisorption energies between those of the same-sized pure titanium and zirconium clusters. The dissociative chemisorption of H2 on a cluster is always accompanied with a charge transfer process. From Table 3, it is clear that the H atoms attract electrons from the host clusters, resulting in the formation of metal hydride. It is found that when the H atoms are located at the bridge sites of the host clusters, more electrons transfer to the two H atoms. For example, for n + m = 5, Ti3Zr2 has transfered more charges to the H atoms when compared with the other clusters. As the host clusters become positively charged after hydrogenations, we can compare the chemisorption energies with the vertical ionization potentials of the clusters. In general, except for Ti3Zr, the trends for ΔECE and VIP are opposite; this is because it is harder for the clusters with larger VIP to loose electrons, so they are less likely to transfer electrons to the H atoms, thus showing lower activity towards H2. This results in the lower chemisorption energies of the clusters with higher VIP.

Fig. 9. (color online) Variations of H2 chemisorption energies as a function of the number of Zr atoms for the TimZrnH2 (n + m ≤ 5) systems.

In Fig. 9, another noteworthy phenomenon is that the maximum of ΔECE is found at the Ti3Zr cluster. To better understand it, we present the HOMO and LUMO distributions of the TimZrnH2 clusters with n + m = 4 in Fig. 10. For these clusters, the delocalization for the frontier orbitals is obvious. The s-d hybridization can be observed between H atom and Ti or Zr atom as well as between the TM atoms in the host clusters. For both orbitals, the states between two H atoms and between each H atom and the host cluster are most dense and delocalized for Ti3ZrH2, that is, the interaction is strongest within the Ti3ZrH2 cluster, which we consider is likely to be the reason behind the markedly large chemisorption energy for Ti3Zr towards H2. This also indicates that Ti3Zr shows high activity with respect to hydrogen, thus promising to be applied as a catalyst in some hydrogenation and dehydrogenation processes.

Fig. 10. (color online) HOMO and LUMO distributions of the TimZrnH2 (n + m = 4) clusters.

We finally discuss the magnetic properties of the TimZrnH2 clusters. To compare the hydrogenated clusters with the bare clusters, their spin magnetic moments are listed together in Table 4. In most cases, the total magnetic moments of these clusters have decreased after chemisorbing H2, and there are four TimZrnH2 clusters of which the magnetic moments have quenched, implying their electrons are all paired. It is interesting to find that for the pure titanium clusters, the magnetic moments remain the same after hydrogenation. This may be because H2 has a weak influence on the pure Ti clusters, which can possibly be reflected by their relatively small ΔECE values.

Table 4.

Magnetic moments of the bare (μ0) and hydrogenated (μ) TimZrn (n + m ≤ 5) clusters.

.
4. Conclusion

The equilibrium geometries, stabilities, and electronic properties of small TimZrn (n + m ≤ 5) clusters were investigated by using density functional calculations combining the PBE functional with the def2-TZVP basis set. The ground state structures of the clusters of the same size are similar to each other. According to the averaged binding energies, the clusters become more stable as the cluster size gets larger and as the number of Zr atoms increases. The clusters consisting of more Ti atoms are usually characterized by larger HOMO–LUMO gaps and VIP as well as smaller VEA, implying their electronic structures are of higher stability. Moreover, the interactions between molecular hydrogen and the stable TimZrn clusters were investigated. Our results show that H2 prefers dissociative chemisorptions onto the bridge or face sites of the considered clusters. We calculated the chemisorption energies of these systems and found the pure Zr clusters are more active towards H2 when compared with the Ti counterparts. The maximum of ΔECE appears at the Ti3ZrH2 cluster, which may be understood by analyzing their HOMO and LUMO states, when strong interactions among the H atoms and the host cluster can be observed. Besides, the total spin magnetic moments of these clusters have decreased or remained the same after hydrogenation except for Ti3Zr2.

Reference
[1] Schmid G 1992 Chem. Rev. 92 1709
[2] Lewis L N 1993 Chem. Rev. 93 2693
[3] Halperin W P 1986 Rev. Mod. Phys. 58 533
[4] Baletto F Ferrando R 2005 Rev. Mod. Phys. 77 371
[5] Ferligoj A 1997 Proc. 19th International Conference Information Technology Interfaces June 17–20, 1997 Pula, Croatia 253
[6] Liu S R Zhai H J Castro M Wang L S 2003 J. Chem. Phys. 118 2108
[7] Sakurai M Watanabe K Sumiyama K Suzuki K 1999 J. Chem. Phys. 111 235
[8] Bouderbala W Boudjahem A G 2014 Physica 454 217
[9] Rodríguez-Kessler P L Ricardo-Chávez J L 2015 Chem. Phys. Lett. 622 34
[10] Javan M B 2015 J. Alloys Compd. 643 56
[11] Sosa-Hernández E M Montejano-Carrizales J M Alvarado-Leyva P G 2016 Eur. Phys. J. 70 208
[12] Li W Y Chen F Y 2014 Chin. Phys. 23 117103
[13] Li T H Wong P C Chang S F Tsai P H Jang J S C Huang J C 2017 Mat. Sci. Eng. 75 1
[14] Oh M Y Kim W G Choe H C Ko Y M 2009 Corros. Eng. Sci. Tec. 8 29
[15] María H A D Juan M C Beleń B B Marcos F G Javier S 2007 Chem. Mater. 19 4283
[16] Visintin A Peretti H A Tori C A Triaca W E 2001 Int. J. Hydrogen Ene. 26 683
[17] Fukunaga T Itoh K Hashi K Aoki K 2002 Appl. Phys. 74 957 Suppl.
[18] Idrobo J C Öğüt S Jellinek J 2005 Phys. Rev. 72 085445
[19] Martin S V Pedro H H T Umapada P Jose F R S Jose I R M Jorge A A 2006 J. Phys. Chem. 110 10274
[20] Medina J Coss de R Tapia A Canto G 2010 Eur. Phys. J. 76 427
[21] Sun H Q Ren Y Wu Z F Xu Z 2015 Comput. Theor. Chem. 74 1062
[22] Turgut B Sakir E Masaru H Shoichi T 2000 J. Phys. 8 223
[23] Wang C C Zhao R N Hang J G 2006 J. Chem. Phys. 124 194301
[24] Zhao W J Lei X L Yan Y L Yang Z Luo Y H 2007 Acta Phys. Sin. 56 5209 in Chinese
[25] Du J G Sun X Y Jiang G 2009 Eur. Phys. J. 55 111
[26] Hurtabo R B Cortez-Valadez M Gámez-Corrales R Flores-Acosta M 2018 Comput. Theor. Chem. 32 1124
[27] Rodríguez-Kessler P L Rodríguez-Domínguez A R 2016 J. Chem. Phys. 120 2401
[28] Zhao G F Sheng X F Zhi L L Sun J M Gu Y Z 2009 J. Mol. Struc. Theochem. 908 40
[29] Chattaraj D Bhattacharya S Dash S Majumder C 2016 J. Appl. Phys. 120 094301
[30] Sheng X F Zhao G F Zhi L L 2008 J. Phys. Chem. 112 17828
[31] Kumar T J D Weck P F Balakrishnan N 2007 J. Phys. Chem. 111 7494
[32] Tarakeshwar P Kumar T J D Balakrishnan N 2008 J. Phys. Chem. 112 2846
[33] Shang M H Wei S H Zhu Y J 2009 J. Phys. Chem. 113 1550
[34] Ge G X Yan H X Jing Q Luo Y H 2011 J. Clust. Sci. 22 473
[35] Ling G 2012 J. Nanopart. Res. 14 957
[36] Zhong M M Huang C Wang G Z 2017 J. Alloys Compd. 725 388
[37] Wen J H Chen G X Zhang J M Wu H 2018 J. Phys. Chem. Solids 115 84
[38] Wang Q Li Q Dai J F Guo Y Q 2012 J. At. Mol. Phys. 29 437
[39] Andrae D Haubermann U Dolg M Stoll H Preub H 1990 Theor. Chim. Acta 77 123
[40] Weigend F Ahlrichs R 2005 Phys. Chem. Chem. Phys. 7 3297
[41] Lombardi J R Davis B 2002 Chem. Revs. 102 2431
[42] Salazar-Villanueva M Hernández Tejeda P H Pal U Rivas-Silva J F Rodríguez Mora J I Ascencio J A 2006 J. Phys. Chem. 110 10274
[43] Lide D R 2007 CRC Handbook Chem. Phys. 88 Boca Ration CRC Press 1024