Next Article in Journal / Special Issue
Elliptical Quantum Rings with Variable Heights and under Spin–Orbit Interactions
Previous Article in Journal / Special Issue
Double Quantum Ring under an Intense Nonresonant Laser Field: Zeeman and Spin-Orbit Interaction Effects
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A DFT + U Study on the Stability of Small CuN Clusters (N = 3–6 Atoms): Calculation of Phonon Frequencies

by
Luis A. Alcalá-Varilla
1,2,3,*,
Rafael E. Ponnefz-Durango
1,
Nicola Seriani
4,
Eduard Araujo-Lopez
5 and
Javier A. Montoya
5
1
Departamento de Física y Electrónica, Universidad de Córdoba, Montería 230002, Córdoba, Colombia
2
Doctorado en Ciencias Físicas, Universidad de Cartagena, Cartagena de Indias 130001, Bolívar, Colombia
3
Departamento de Ingeniería de Sistemas, Universidad Cooperativa de Colombia, Montería 230002, Córdoba, Colombia
4
Condensed Matter and Statistical Physics Section, The Abdus Salam ICTP, Strada Costiera 11, 34151 Trieste, Italy
5
Grupo de Modelado Computacional, Universidad de Cartagena, Cartagena de Indias 130001, Bolívar, Colombia
*
Author to whom correspondence should be addressed.
Condens. Matter 2023, 8(3), 81; https://doi.org/10.3390/condmat8030081
Submission received: 29 June 2023 / Revised: 28 August 2023 / Accepted: 5 September 2023 / Published: 11 September 2023
(This article belongs to the Special Issue Physics of Light-Matter Coupling in Nanostructures)

Abstract

:
Despite the interest in copper clusters, a consensus on their atomic structure is still lacking. The experimental observation of isolated clusters is difficult, and theoretical predictions vary widely. The latter is because one must adequately describe the closed shell of d electrons both in its short- and long-range effects. Herein, we investigate the stability of small copper clusters (Cu N , N = 3–6 atoms) using spin-polarized DFT calculations under the GGA approximation, the Hubbard U correction, and the van der Waals forces. We found that the spin-polarized and vdW contributions have little effect on the binding energies of the isomers. The inclusion of U represents the most relevant contribution to the ordering of the Cu N isomers, and our calculated binding energies for the clusters agreed with the experimental values. We also found that atomic relaxations alone are not enough to determine the stability of small copper clusters. It is also necessary to build the energy landscape or calculate the vibrational frequencies of the isomers. We found that the vibrational frequencies of the isomers were in the THz range and the normal modes of vibration were discrete. This approach is relevant to future studies involving isolated or supported copper clusters.

1. Introduction

Many studies have shown that copper clusters (Cu N ) can be used for CO 2 reduction, for photocatalysis processes [1,2,3,4,5,6], and as dopants to improve the photocatalytic activity of other photocatalysts [7,8,9,10,11,12,13,14,15]. Due to the importance of copper clusters, there are many theoretical works that have studied these structures. For example, the structure and stability of small copper clusters were studied using DFT by Jug et al. in 2002 [16]. The energy and structure of copper clusters were studied using the Monte Carlo method by Zhang et al. in 2007 [17]. The structures and electronic properties of copper clusters were studied using DFT by Cui-Ju et al. in 2009 [18]. A study of the structure and an analysis of the atomic vibrations of copper clusters were carried out by Rusina et al. in 2013 [19], whereas a density functional study on the stability and structural properties of Cu N clusters was presented by Ketabi et al. in 2013 [20]. The atomic structures of small Cu clusters with 3–6 atoms were investigated using density functional theory and a random search algorithm by Cogollo-Olivo in 2015 [21]. Recently (2020), Ahmed reported a study in which the structural properties of Cu N clusters (N = 1–7) were determined [22]. Experimental studies have also been reported [23,24,25,26,27]. Finally, studies can be found in which the properties of small copper clusters are compared with those of copper surfaces [28,29].
At this point, it is clear that copper clusters have great relevance and very promising applications. This is reason enough to review their current research status. We have found that small copper clusters are frequently added to surfaces to improve their properties. The interaction of Cu N cluster (N = 1–4) nanoparticles with ChCl:Urea deep eutectic solvent was studied by Ghenaatian and coworkers in 2021 [12], whereas the stability of Cu N clusters (N = 1–4) adsorbed on CuAlO 2 surfaces was studied using atomic thermodynamics by Wang and coworkers in 2022 [14]. The frequent use of these sizes of Cu N clusters motivated us to review the structures reported as stable for small copper clusters in vacuum, finding large differences between the results reported in previous works, despite the fact that some of them used the same level of theory [16,21,22,30,31,32,33,34,35,36,37,38,39,40,41]. These differences further motivated us to carry out a study on the stability of small copper clusters, considering the frequently used cluster sizes (3–6 atoms), where the greatest discrepancies occur. With this work, we hope to provide a better understanding of the structural properties of small copper clusters, as well as clarify the differences between the results previously reported in other works.
Also, from a fundamental standpoint, by revisiting the structural properties of small isolated Cu N clusters with a more accurate level of theory compared to most previous works, our results provide a stronger foundation for future studies that require good knowledge of the systematic reasons that can explain the relative stabilities of these forms. Here, an improved description of the clusters is achieved, for instance, by sampling a wide range of values of the Hubbard U and gradually observing their effects. On the other hand, although the impact of the van der Waals interactions on the structural properties of copper clusters seems to be minor, judging from our results, we observe that it still may become important in future studies and does not negatively affect the description of these structures obtained using DFT + U.
After this section, the rest of this work is organized as follows. In Section 2, we present the computational details necessary to reproduce all the results in this work. A discussion is also provided in the framework of DFT on the different levels of theory that we consider necessary to study the stability of small copper clusters. Section 3 is divided into four main parts. First, we present the Cu N clusters (N = 3–6) that were optimized in this work using DFT + GGA (ground states and other metastable isomers), and we discuss the differences between our results and others reported in the literature. Binding energies are then calculated for all structures using different levels of theory, including spin-polarized, Hubbard U correction, and van der Waals dispersion forces. In this part, we can see how the energy ordering of the isomers changes with the inclusion of the U term and also how the energy landscape of Cu 3 changes with the inclusion of this term. Binding energies are also calculated using the HSE hybrid functional, and all results are compared with those of other experimental and theoretical works. Since it is not easy to build the energy landscape for Cu 4 , Cu 5 , and Cu 6 to identify points of minimum energy, maximum energy, and saddle points for these structures, we calculate the densities of the state of phonons so that we can determine which of the structures found for Cu 4 , Cu 5 , and Cu 6 are really stable. In the last part of Section 3, the structural and electronic properties of Cu N copper clusters are shown. Finally, we present the conclusions of this work in Section 4.

2. Materials and Methods

We started by reproducing a complete set of the most stable copper-cluster isomers (Cu N for N = 3–6 atoms) that have been reported in the literature [21,36], while also performing a heuristic structural search. After this, ab initio atomic relaxations were performed for each of the studied cases independently, using three different levels of theory (SGGA, SGGA + U, and SGGA + U + vdW). The Hubbard term correction was applied using the simplified version of Dudarev et al. [42], and the vdW dispersion forces were included using Grimme’s DFT + D2 approach [43].
All the DFT calculations were performed with the Quantum ESPRESSO [44,45] package, using the Perdew–Burke–Ernzerhof (PBE) generalized gradient (GGA) exchange-correlation functionals [46], which was generated using a scalar-relativistic calculation and the nonlinear core correction. Vanderbilt ultra-soft pseudopotentials were employed [47], and a plane-wave basis set was used. LSDA spin-polarized calculations were taken into account. The tested values of U ranged from 0 to 9 eV. The convergence of the total energy for the clusters was achieved using cutoffs of 60 and 600 Ry for the kinetic energy of the wavefunctions and the charge density, respectively. A grid of 2 × 2 × 2 k-points was generated using the method of Monkhorst and Pack [48], which was selected based on the convergence of the total energy with respect to the k-points.
To avoid interactions between neighboring periodic images, copper clusters were placed inside a large cubic cell with an edge of 14 Å. The converged cell size was determined by analyzing the stabilization within 5 meV/atom of the total energy with respect to the cell parameters for the largest isomer included in this study (Cu 6 ), taken as the worst-case scenario. Atomic relaxations were performed for each increase in cell size using the Broyden–Fletcher–Goldfarb–Shanno (BFGS) algorithm until the forces and energies converged within 1 × 10 4 and 1 × 10 6 a.u., respectively. The average binding energy was defined as
B E = n × E C u E C u n / n
where E ( C u ) is the electronic energy of a Cu atom and E ( C u n ) is the electronic energy of the cluster of n atoms.
We now discuss why it is important to take into account the Hubbard U correction when studying the stability of small copper clusters. The peculiar electronic structure ( 3 s 2 3 p 6 3 d 10 4 s 1 ) of copper leads to a subtle interplay between the one s electron and the closed d shell, resulting in complex behavior, which is difficult to capture using theoretical methods. On the other hand, Density Functional Theory (DFT) is a convenient theoretical tool that is frequently used to characterize the properties of photocatalysts with impurities. It is well-known that basic descriptions of the exchange-correlation functionals, such as the local density and generalized gradient approximations (LDA and GGA, respectively), fail to predict the presence of some states in the electronic structure of the supported Cu N /photocatalyst system and also overestimate the binding energies for molecules and clusters [8,49,50,51,52]. This failure is due to the self-interaction error inherent in these exchange-correlation functionals, leading to small energy differences between the atomic s and d levels and strong hybridization effects [53]. To correct this error, post-DFT methods, such as DFT + U and hybrid functionals, have been employed [8,49,50,51,52,54,55,56,57,58,59,60,61,62,63]. However, hybrid approaches result in greater computational costs, which turns out to be critical for calculations such as those involving supported clusters via the supercell method to describe rare events. On the other hand, the Strongly Constrained and Appropriately Normed (SCAN) semilocal density functional has been shown to be a great alternative in the DFT framework [64]; however, Tameh and coworkers [65] showed that the Gibbs energies and the binding energies for the hydrogenation of CO and CO 2 on Cu(211) calculated using the SCAN functional are worse than those obtained using PBE and HSE. Moreover, Long and coworkers recently evaluated the optimal U value for 3 d transition-metal oxides within the SCAN + U framework [66] and found that using DFT-SCAN, the variation of the oxidation reaction enthalpy overestimates the experimental enthalpy by ∼0.7 eV, and when the U term is included, it monotonically worsens the error between theoretical (SCAN + U) and experimental enthalpies.
We also provide an explanation of the inclusion of the van der Waals dispersion forces in this work. It is known that vdW is a long-range interaction, and, therefore, this force is really important in adsorption studies, such as contaminant adsorption on copper clusters or copper clusters’ adsorption on surfaces. Therefore, we have included this interaction to provide information on the minor effects of it on the stability of isolated copper clusters that can be used in future studies. Additionally, the presence of a full d shell not only affects the short-range electron–electron interaction but also influences long-range electronic correlations (van der Waals interactions). It is then clear that performing calculations accurately in both ranges is extremely important. However, although several theoretical works based on DFT have studied the structure of small copper clusters [16,21,22,30,31,32,33,34,35,36,37,38,39,40,41], to the best of our knowledge, no previous work has included the Hubbard U correction and van der Waals (vdW) dispersion forces simultaneously within a spin-polarized treatment (SGGA + vdW + U for short). By the end, we hope to convey the message that this level of treatment is indeed relevant, with important implications for future studies involving interactions between supported Cu clusters and molecules, where adsorbate-metal hybridization interaction, as well as physisorption governed by dispersion forces, are frequently present.

3. Results and Discussion

This section begins with the presentation of the Cu N clusters (N = 3–6 atoms), which were obtained in this work using DFT-GGA and atomic relaxations. These isomers are shown in Table 1, which shows two structures for Cu 3 , four for Cu 4 , five for Cu 5 , and seven for Cu 6 .
Although the atomic relaxations suggest that all the structures in Table 1 are stable, we will demonstrate that some of them are not, thereby showing that DFT-GGA alone is not sufficient to predict the correct structures of small copper clusters. On the other hand, based on our current knowledge, previous theoretical studies by other authors have only considered the GGA approach in the DFT framework to study the structures of small copper clusters, and, therefore, we can find different results among similar works. For example, Cogollo-Olivo et al. [21] and Ahmed et al. [22] reported two stable isomers for Cu 3 , one of them being an isosceles triangle (GS) and the other a bent-type isomer. Moreover, Guvelioglu et al. [36] also used DFT-GGA and reported two stable triangular isomers for Cu 3 , whereas Lammers and Borstel [67] used the tight-binding method and instead reported four. In our case, the DFT-GGA atomic relaxations suggest an equilateral triangle as the GS of Cu 3 (A isomer in Table 1) and a linear isomer as the stable structure (B isomer in Table 1). An equilateral triangle was also proposed as the GS for Cu 3 in [18,19], and a linear isomer was also proposed as the stable GS for Cu 3 in [18]. In the same way, for Cu 4 , Guvelioglu et al. [36] and Cui-Ju et al. [18] suggested the possibility of a metastable tetrahedral structure (using DFT-GGA), but Cogollo-Olivo et al. [21], using the same level of theory, showed that this isomer is not stable, whereas our DFT-GGA atomic relaxations also predict the possibility of a tetrahedral structure (F isomer in Table 1), Furthermore, this tetrahedral isomer was proposed as the GS for Cu 4 in [17,19] who used the Monte Carlo and tight-binding methods, respectively. Structures C and D in Table 1 are in agreement with those reported in [21,22,36], whereas the E isomer for Cu 4 was also reported in [36]. Our atomic relaxations do not predict the other Cu 4 isomers reported in [21]. Regarding Cu 5 (clusters with this particular size have been already synthesized and proposed as promising catalysts [1,5]), the G and H isomers in Table 1 were also reported by Cogollo-Olivo et al. [21], Ahmed et al. [22], and Guvelioglu et al. [36], whereas our DFT-GGA atomic relaxations also predict the I, J, and K structures for Cu 5 . Moreover, the J isomer was also suggested as stable by the authors of [18], and the H structure was proposed as the GS for Cu 5 in [17]. Finally, for Cu 6 , the L, M, N, and O structures in Table 1 were also reported by Cogollo-Olivo et al. [21], whereas the P and Q isomers were reported in [36], and the last structure (R) for Cu 6 was obtained by our DFT-GGA atomic relaxations. In addition, the L isomer was also proposed by the authors of [18,20,22], the M, N, and P isomers were also proposed by the authors of [18], and the P isomer was also proposed as the GS for Cu 6 by the authors of [17,19], who used the Monte Carlo and tight-binding methods, respectively. To clarify why there are some different results in the works mentioned above and identify which structures in Table 1 are really stable, we use other approaches in the next subsections of this work.

3.1. Binding Energies of Cu N Clusters

Binding energies ( B E ) are usually used to determine the ground state (GS) of each group of isomers. The isomer with the highest binding energy is the GS, and the isomers with negative binding energies are not stable. In this work, the binding energies were calculated for all the isomers in Table 1 using DFT + SGGA + vdW + U (DFT with GGA, spin-polarized calculations, van der Waals (vdW) dispersion forces, and Hubbard-U correction), the values of U term were varied from 0 to 9 eV, and the results are shown in Table 2, Table 3 and Table 4. In these tables, the isomers A, C, G, and L are the GS of each group of copper clusters with 3–6 atoms, respectively. In Table 2, Table 3 and Table 4, Δ B E are the relative binding energies and the binding energy differences between each isomer and the corresponding GS.
In this study, spin-polarized calculations and vdW forces have minor relevance in determining the binding energies’ values, whereas the U term produces larger changes in the B E values (see Table 2, Table 3 and Table 4). B E decreases for all structures in Table 1 when the value of U increases. This is the first effect we can observe due to the inclusion of U. The second is that the decrease in B E is not in the same proportion for all isomers. In some, the decrease is greater than in others, which can be observed in the data for Δ B E in Table 2, Table 3 and Table 4. For some structures, Δ B E increases, and for others, it decreases when the U term enhances. Therefore the U term produces an energetic rearrangement in the structures. Note that when U = 0 , the energetic arrangement of the isomers is as shown in Table 5.
For U = 8 eV, the energetic arrangement of the isomers is as shown in Table 6. In Table 5 and Table 6, it can be seen that the GS of the structures is the same for U = 0 and U = 8 eV, but for Cu 5 and Cu 6 , there are changes in the order of some isomers. For example, when U = 0 , the second isomer of Cu 5 is (H), the third is (I), the fourth is (J), and the fifth is (K). But when U = 8 eV, the (I) isomer becomes the second, the (K) isomer is now the third, and H and J become the fourth and fifth, respectively. Therefore, we can say that the U term favors 2D structures because the Δ B E of the 3D structures (H and J) increases more than the Δ B E of the planar isomers (I and K). A similar analysis can be made for Cu 6 , where for U = 8 eV, the O, N, P, R, and Q isomers become the third, fourth, fifth, sixth, and seventh, respectively. In addition, for U = 9 eV, the N isomer of Cu 6 does not converge. But for U < 9 eV, that structure does converge. So the following question can be asked: What is the correct value of U for copper clusters? We provide a response to that question later.
On the other hand, recalling the main objective of this work, it can be observed in Table 2, Table 3 and Table 4 that the atomic relaxations with the U term can still predict stable copper clusters. We proceed to carry out another type of analysis. For this, we start with Cu 3 clusters.

3.1.1. Energy Landscape for Cu 3 Copper Clusters

Figure 1 shows a general Cu 3 copper cluster. It is possible to find all the isomers for Cu 3 by varying the angle ( θ ) and calculating the optimal interatomic distance (d).
In Figure 2, the energy landscape of Cu 3 is shown. To make this graph, we fixed several angles (from 0 to 180 degrees in steps of 5 degrees) and identified the optimal d distance, which minimizes the energy for each of them. We see that using just GGA (black curve), it predicts two minimum energy points ( θ = 60 ° , 118 ° ), representing two stable isomers, as reported in [21,22]. The minimum energy point (near 118 ° ) tends to disappear when the spin-polarized calculations (red curve) and vdW dispersion forces (green curve) are introduced (here, the first effect due to spin-polarized calculations and vdW forces is observed). Finally, when the Hubbard U term is taken into account (blue curve), there is only one minimum energy point on the energy landscape of Cu 3 , corresponding to one stable isomer for Cu 3 , which is an equilateral triangle ( θ = 60 ° ). The inclusion of the term U rules out the possibility of finding the metastable isosceles triangle reported in previous DFT studies [21,22,36]. On the other hand, in Figure 2, it can be seen that the linear isomer for Cu 3 corresponds to a maximum energy point ( θ = 180 ° ); therefore, the linear isomer Cu 3 (B) in Table 1 is unstable. By analyzing the behavior of the Cu 3 (B) isomer, we can see that the atomic relaxations could converge on structures that correspond to points of maximum energy in the energy landscape. However, the point of maximum energy seen in Figure 2 at θ = 180 ° is really a saddle point. In other words, at this point, the energy is a function of two variables E( θ , d). Therefore, for variations of θ , the behavior of the energy is as shown in Figure 2, whereas for a fixed θ value ( θ = 180 ° ) and variations of the interatomic distance (d), E( 180 ° , d) has a minimum at θ = 180 ° . So we can ask: Which of the structures in Table 1 are also unstable (saddle points for the energy)? Building the energy landscape for Cu 4 , Cu 5 , and Cu 6 is not easy, so we must find other ways to determine the stability of copper clusters.

3.1.2. Binding Energies Calculated Using the HSE Hybrid Functional

Using the HSE hybrid functional, atomic relaxations were also performed for all structures in Table 1, and the values of the binding energies are shown in Table 2, Table 3 and Table 4. It can be seen that the B E values calculated using the HSE are lower than those found using SGGA + vdW + U. On the other hand, it can be observed that the HSE indicates that the isomers Cu 4 (F), Cu 5 (J), and Cu 6 (P) are not stable, in contrast to the predictions using SGGA + vdW + U. Table 2, Table 3 and Table 4 also show that the HSE indicates a stable configuration for Cu 6 (N) in accordance with the SGGA + vdW + U predictions for this structure for U < 9 eV. The latter is very interesting because SGGA + vdW + U predicts that Cu 6 (N) is unstable for U = 9 eV, so we may think that there should be a limit for taking a correct value of U for copper clusters. In fact, this is confirmed by looking at Figure 3a, where graphs of the B E values are shown as a function of the size of the copper cluster (for the GS) for different levels of theory.
In Figure 3a, it can be seen that when the value of U increases, the B E values tend to approach the experimental values reported by Spasov et al. (2000) [23]. It can also be observed that the B E values calculated using the HSE hybrid functional are further from the experimental data than those calculated using SGGA + vdW + U. On the other hand, Table 2, Table 3 and Table 4 also show the values of the binding energies reported by other authors [17,18,19,20,21,22,23,36], and Figure 3b shows a comparison of the binding energies calculated in this work for U = 5 and U = 8 eV with those reported in other works. It can be seen that our results are closer to the experimental values than those previously calculated, even for U = 5 eV. Therefore, we can suggest any value of U in the interval ( 5 U 8 ) eV as a suitable choice to work with copper atoms. These values of U are, in any case, in line with those of previous works that have used values for U between 4 and 8 eV for copper and copper oxides [8,68,69,70,71,72]. In addition, in Figure 3a, it can be seen that the values of the binding energies calculated using only spin polarization and vdW (sp-vdW) are closer to the experimental data compared to those found using only PBE. This shows the importance of including these interactions in this work.
Remembering again that the main objective of this work is to study the stability of the structures in Table 1, the only thing we can say up until now is that for Cu 3 , there is only one stable isomer. For Cu 4 , Cu 5 , and Cu 6 , we still cannot ensure that these isomers are stable because we did not build their energy landscapes, and as we mentioned before, this is not easy. Therefore, we choose to calculate the vibration frequencies of these structures to investigate which of them are really stable.

3.2. Density of States of Phonons of Cu N Copper Clusters

To finish the investigation on the stability of the copper clusters in Table 1, we calculated the vibration frequencies of these isomers using SGGA + vdW + U. At this point, we know from the previous subsection what values of U we can use, so we used U = 8 eV to continue. In Figure 4, the phonon state densities for U = 8 eV are presented. Note that Cu 3 (B) has negative frequencies, confirming that this isomer is unstable. It can also be seen that the phonon state densities of the isomers Cu 4 (D), Cu 4 (F), Cu 5 (I), Cu 5 (J), Cu 5 (K), and Cu 6 (Q) also have negative frequencies; therefore, those isomers are also unstable. Remember that the HSE hybrid functional also predicted that Cu 4 (F) and Cu 5 (J) were unstable. Furthermore, the HSE suggested that Cu 6 (P) was unstable, but the atomic relaxation using SGGA + vdW + U and the phonon state density predicted that Cu 6 (P) was stable. In this case, we think that Cu 6 (P) is really stable, considering that the results of the B E values using SGGA + vdW + U were closer to the experimental data compared to those obtained using the HSE (see Figure 3a).
Of the 18 initial structures in Table 1, only 11 are stable: one for Cu 3 , two for Cu 4 , two for Cu 5 , and six for Cu 6 . We have seen that only the atomic relaxations using SGGA + vdW + U or the HSE hybrid functional were not enough to determine the stability of the small copper clusters in this study. We also needed to calculate the energy landscape (as for Cu 3 ) or the vibrational frequencies of the isomers to determine their stability.
As an additional result, in Figure 4, it can be seen that the vibrational frequencies of the copper clusters are between 0 and 10 THz.

3.3. Structural and Electronic Properties of Cu N Copper Clusters

In this subsection, the main structural and electronic properties of the eleven stable copper clusters are presented. Table 7, Table 8 and Table 9 show the bond lengths of the isomers for three different values of the U term. It can be seen that when the U term increases, all the bond lengths of the isomers also increase. This helps to explain why the binding energies of the isomers decrease when the U term increases. We also calculated de density of states (DOS) of these isomers (see Figure 5) and found that when the U term increases, an intragap appears within the valence band (see the intragap values in Table 7 and Table 9). The latter is also associated with the decrease in the binding energies of the isomers when the U term increases. The greatest intragaps occur for 3D structures such as Cu 5 (H), Cu 6 (N), and Cu 6 (P), so by increasing the U term, they have a greater decrease in the binding energies (see Table 2, Table 3 and Table 4).
On the other hand, regardless of the value of U, we found the following property for the isomers in our study: the binding energies increase with the cluster size. The ground state of the Cu 3 and Cu 6 isomers exhibit a D 3 h symmetry, whereas the GS for the Cu 4 and Cu 5 structures exhibit D 2 h and C 2 v symmetries, respectively. The Cu 3 and Cu 5 copper clusters exhibit a magnetization close to 1 a.u. due to the polarization of their atoms (see Table 7). Isomers with an even number of atoms exhibit no magnetization, but there is an interesting exception for the Cu 6 (P) and Cu 6 (R) isomers, which exhibit a magnetization of around 2 a.u., also due to the polarization of their atoms (see Table 9). In Figure 5, it can be seen that the isomers with magnetization do not exhibit a symmetric DOS around the Fermi level where the 3 p and 4 s orbitals of the copper atoms predominate.

4. Conclusions

We have carried out a study on the stability of small copper clusters (Cu N , N = 3–6 atoms) using DFT and have found that the GGA approximation alone is insufficient to predict the stability of these structures or the correct ordering of the binding energies of these isomers. The Hubbard U correction is absolutely necessary to predict the stability and correct ordering in the binding energies of small copper clusters. Based on our results, we think that this correction should always be taken into account in any study involving copper atoms. We recommend working with a value of U between 5 and 8 eV. We have also found that atomic relaxations alone are insufficient to determine the stability of clusters. Atomic relaxations based on hybrid functionals such as the HSE could rule out some structures, but they are insufficient since the isomers obtained after the relaxations could be associated with energy saddle points. To determine whether the isomers obtained with atomic relaxations are stable, it is necessary to build the energy landscape, as done for Cu 3 in this study, or perform additional calculations. For this, we recommend calculating the frequencies of the phonons. On the other hand, the vibrational frequencies of small copper clusters are of the order of THz (0 to 10 THz), and the normal modes of vibration are discrete.

Author Contributions

L.A.A.-V.: calculations, conceptualization, methodology, formal analysis, investigation, and writing; R.E.P.-D.: methodology and calculations; N.S. and J.A.M.: formal analysis, investigation, supervision, and writing; E.A.-L.: methodology and writing. All authors have read and agreed to the published version of the manuscript.

Funding

The authors would like to thank the ICTP for providing access to its computational resources and the Fondo Nacional de Financiamiento para la Ciencia, la Tecnología y la Innovación “Francisco José de Caldas” through contract 299-2016. N.S. would also like to thank ICETEX for their support through the “Profesores Invitados” program. L.A.A.-V. acknowledges the ICTP as the recipient of a TRIL Programme fellowship. Finally, J.A.M. and L.A.A.-V. would also like to thank the Vicerrectorías de Investigaciones of the Universidades de Cartagena y Córdoba for their support through internal research projects.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Concepción, P.; Boronat, M.; García-García, S.; Fernández, E.; Corma, A. Enhanced Stability of Cu Clusters of Low Atomicity against Oxidation. Effect on the Catalytic Redox Process. ACS Catal. 2017, 7, 3560–3568. [Google Scholar] [CrossRef]
  2. Iyemperumal, S.K.; Deskins, N.A. Activation of CO2 by supported Cu clusters. Phys. Chem. Chem. Phys. 2017, 19, 28788–28807. [Google Scholar] [CrossRef] [PubMed]
  3. Yang, B.; Liu, C.; Halder, A.; Tyo, E.C.; Martinson, A.B.F.; Seifert, S.; Zapol, P.; Curtiss, L.A.; Vajda, S. Copper Cluster Size Effect in Methanol Synthesis from CO2. J. Phys. Chem. C 2017, 121, 10406–10412. [Google Scholar] [CrossRef]
  4. Barabás, J.; Höltzl, T. Reaction of N2O and CO Catalyzed with Small Copper Clusters: Mechanism and Design. J. Phys. Chem. A 2016, 120, 8862–8870. [Google Scholar] [CrossRef] [PubMed]
  5. Huseyinova, S.; Blanco, J.; Requejo, F.G.; Ramallo-López, J.M.; Blanco, M.C.; Buceta, D.; López-Quintela, M.A. Synthesis of Highly Stable Surfactant-free Cu5 Clusters in Water. J. Phys. Chem. C 2016, 120, 15902–15908. [Google Scholar] [CrossRef]
  6. Reske, R.; Mistry, H.; Behafarid, F.; Cuenya, B.R.; Strasser, P. Particle Size Effects in the Catalytic Electroreduction of CO2 on Cu Nanoparticles. J. Am. Chem. Soc. 2014, 136, 6978–6986. [Google Scholar] [CrossRef] [PubMed]
  7. Schlexer, P.; Chen, H.-Y.T.; Pacchioni, G. CO2 Activation and Hydrogenation: A Comparative DFT Study of Ru10/TiO2 and Cu10/TiO2 Model Catalysts. Catal. Lett. 2017, 147, 1871–1881. [Google Scholar] [CrossRef]
  8. Seriani, N.; Pinilla, C.; Crespo, Y. Presence of Gap States at Cu/TiO2 Anatase Surfaces: Consequences for the Photocatalytic Activity. J. Phys. Chem. C 2015, 119, 6696–6702. [Google Scholar] [CrossRef]
  9. Roy, S.C.; Varghese, O.K.; Paulose, M.; Grimes, C.A. Toward Solar Fuels: Photocatalytic Conversion of Carbon Dioxide to Hydrocarbons. ACS Nano 2010, 4, 1259–1278. [Google Scholar] [CrossRef]
  10. Varghese, O.K.; Paulose, M.; LaTempa, T.J.; Grimes, C.A. High-Rate Solar Photocatalytic Conversion of CO2 and Water Vapor to Hydrocarbon Fuels. Nano Lett. 2009, 9, 731–737. [Google Scholar] [CrossRef]
  11. Shankar, K.; Basham, J.I.; Allam, N.K.; Varghese, O.K.; Mor, G.K.; Feng, X.; Paulose, M.; Seabold, J.A.; Choi, K.-S.; Grimes, C.A. Recent Advances in the Use of TiO2 Nanotube and Nanowire Arrays for Oxidative Photoelectrochemistry. J. Phys. Chem. C 2009, 113, 6327–6359. [Google Scholar] [CrossRef]
  12. Ghenaatian, R.H.; Shakourian-Fard, M.; Kamath, G. Interaction of Cun, Agn and Aun (n = 1–4) nanoparticles with ChCl: Urea deep eutectic solvent. J. Mol. Graph. Model. 2021, 105, 107866. [Google Scholar] [CrossRef] [PubMed]
  13. Zhang, R.; Chutia, A.; Sokol, A.A.; Chadwick, D.; Catlow, C.R.A. A computational investigation of the adsorption of small copper clusters on the CeO2 (110) surface. Phys. Chem. Chem. Phys. 2021, 23, 19329–19342. [Google Scholar] [CrossRef] [PubMed]
  14. Wang, H.; Wang, L.; Cheng, L.; Feng, G.; Yu, X. Stability of Cun (n = 1–4) clusters adsorption on CuAlO2 ( 11 2 ¯ 0 ) surfaces using atomic thermodynamics. Appl. Surf. Sci. 2022, 578, 151935. [Google Scholar] [CrossRef]
  15. Huang, Z.; Zuo, X.; Li, D.; Qu, B.; Zhou, R. Evolution of atomic and electronic structures of Cun (n = 2–10) clusters supported on anatase TiO2 (101) surface. Appl. Surf. Sci. 2022, 576, 151772. [Google Scholar] [CrossRef]
  16. Jug, K.; Zimmermann, B.; Calaminici, P.; Köster, A.M. Structure and stability of small copper clusters. J. Chem. Phys. 2002, 116, 4497–4507. [Google Scholar] [CrossRef]
  17. Zhang, M.L.; Li, G.P. Energy and Structure of Copper Clusters (n = 2–70,147,500) Studied by the Monte Carlo Method. Solid State Phenom. 2007, 121, 607–610. [Google Scholar] [CrossRef]
  18. Feng, C.-J.; Zhang, X.-Y. Structures and Electronic Properties of CuN (N ≤ 13) Clusters. Commun. Theor. Phys. 2009, 52, 675. [Google Scholar]
  19. Rusina, G.G.; Borisova, S.D.; Chulkov, E.V. Structure and analysis of atomic vibrations in clusters of Cun (n ≤ 20). Russ. J. Phys. Chem. A 2013, 87, 233–239. [Google Scholar] [CrossRef]
  20. Ketabi, S.; Ghasemi, G. Density Functional Study on Stability and Structural Properties of Cun clusters. J. Phys. Theor. Chem. IAU Iran 2013, 9, 223–230. [Google Scholar]
  21. Cogollo-Olivo, B.H.; Seriani, N.; Montoya, J.A. Unbiased structural search of small copper clusters within DFT. Chem. Phys. 2015, 461, 20–24. [Google Scholar] [CrossRef]
  22. Ahmed, A.A. Structural and electronic properties of the adsorption of nitric oxide molecule on copper clusters CuN (N = 1–7): A DFT study. Chem. Phys. Lett. 2020, 753, 137543. [Google Scholar] [CrossRef]
  23. Spasov, V.A.; Lee, T.-H.; Ervin, K.M. Threshold collision-induced dissociation of anionic copper clusters and copper cluster monocarbonyls. J. Chem. Phys. 2000, 112, 1713–1720. [Google Scholar] [CrossRef]
  24. Vogel, M.; Herlert, A.; Schweikhard, L. Photodissociation of small group-11 metal cluster ions: Fragmentation pathways and photoabsorption cross sections. J. Am. Soc. Mass Spectrom 2003, 14, 614–621. [Google Scholar] [CrossRef] [PubMed]
  25. Lecoultre, S.; Rydlo, A.; Félix, C.; Buttet, J.; Gilb, S.; Harbich, W. Optical absorption of small copper clusters in neon: Cun, (n = 1–9). J. Chem. Phys. 2011, 134, 074303. [Google Scholar] [CrossRef] [PubMed]
  26. Pettiette, C.L.; Yang, S.H.; Craycraft, M.J.; Conceicao, J.; Laaksonen, R.T.; Cheshnovsky, O.; Smalley, R.E. Ultraviolet photoelectron spectroscopy of copper clusters. J. Chem. Phys. 1988, 88, 5377–5382. [Google Scholar] [CrossRef]
  27. Ho, J.; Ervin, K.M.; Lineberger, W.C. Photoelectron spectroscopy of metal cluster anions: Cun, Agn, and Aun. J. Chem. Phys. 1990, 92, 6987–7002. [Google Scholar] [CrossRef]
  28. Crispin, X.; Bureau, C.; Geskin, V.; Lazzaroni, R.; Bŕedas, J.L. Local density functional study of copper clusters: A comparison between real clusters, model surface clusters, and the actual metal surface. Eur. J. Inorg. Chem. 1999, 1999, 349–360. [Google Scholar] [CrossRef]
  29. Schmid, G. Large clusters and colloids. Metals in the embryonic state. Chem. Rev. 1992, 92, 1709–1727. [Google Scholar] [CrossRef]
  30. Fernández, E.; Boronat, M.; Corma, A. Trends in the Reactivity of Molecular O2 with Copper Clusters: Influence of Size and Shape. J. Phys. Chem. C 2015, 119, 19832–19846. [Google Scholar] [CrossRef]
  31. Jaque, P.; Toro-Labbé, A. Polarizability of neutral copper clusters. J. Mol. Model. 2014, 20, 2410. [Google Scholar] [CrossRef] [PubMed]
  32. Afshar, M.; Sargolzaei, M. Spin and orbital magnetism of coinage metal trimers (Cu3, Ag3, Au3): A relativistic density functional theory study. AIP Adv. 2013, 3, 112122. [Google Scholar] [CrossRef]
  33. Ren, X.; Rinke, P.; Scuseria, G.E.; Scheffler, M. Renormalized second-order perturbation theory for the electron correlation energy: Concept, implementation, and benchmarks. Phys. Rev. B 2013, 88, 035120. [Google Scholar] [CrossRef]
  34. Baishya, K.; Idrobo, J.C.; Ögüt, S.; Yang, M.; Jackson, K.A.; Jellinek, J. First-principles absorption spectra of Cun (n = 2–20) clusters. Phys. Rev. B 2011, 83, 245402. [Google Scholar] [CrossRef]
  35. Böyükata, M.; Belchior, J.C. Structural and energetic analysis of copper clusters: MD study of Cun (n = 2–45). J. Braz. Chem. Soc. 2008, 19, 884–893. [Google Scholar] [CrossRef]
  36. Guvelioglu, G.H.; Ma, P.; He, X.; Forrey, R.C.; Cheng, H. First principles studies on the growth of small Cu clusters and the dissociative chemisorption of H2. Phys. Rev. B 2006, 73, 155436. [Google Scholar] [CrossRef]
  37. Li, S.; Alemany, M.M.G.; Chelikowsky, J.R. Real space pseudopotential calculations for copper clusters. J. Chem. Phys. 2006, 125, 034311. [Google Scholar] [CrossRef] [PubMed]
  38. Yang, M.; Jackson, K.A. First-principles investigations of the polarizability of small-sized and intermediate-sized copper clusters. J. Chem. Phys. 2005, 122, 184317. [Google Scholar] [CrossRef]
  39. Kabir, M.; Mookerjee, A.; Bhattacharya, A.K. Structure and stability of copper clusters: A tight-binding molecular dynamics study. Phys. Rev. A 2004, 69, 043203. [Google Scholar] [CrossRef]
  40. Massobrio, C.; Pasquarello, A.; Car, R. Structural and electronic properties of small copper clusters: A first principles study. Chem. Phys. Lett. 1995, 238, 215–221. [Google Scholar] [CrossRef]
  41. Ramírez-Solís, A.; Novaro, O.; Ruiz, M.E. Stability of Cu3: A variational and perturbational configuration-interaction study. Phys. Rev. B 1987, 35, 4082–4084. [Google Scholar] [CrossRef] [PubMed]
  42. Dudarev, S.L.; Botton, G.A.; Savrasov, S.Y.; Humphreys, C.J.; Sutton, P. Electron-energy-loss spectra and the structural stability of nickel oxide: An LSDA + U study. Phys. Rev. B 1998, 57, 1505–1509. [Google Scholar] [CrossRef]
  43. Grimme, S. Semiempirical GGA-type density functional constructed with a long-range dispersion correction. J. Comput. Chem. 2006, 27, 1787–1799. [Google Scholar] [CrossRef] [PubMed]
  44. Giannozzi, P.; Baroni, S.; Bonini, N.; Calandra, M.; Car, R.; Cavazzoni, C.; Ceresoli, D.; Chiarotti, G.L.; Cococcioni, M.; Dabo, I.; et al. QUANTUM ESPRESSO: A modular and open-source software project for quantum simulations of materials. J. Phys. Condens. Matter 2009, 21, 395502. [Google Scholar] [CrossRef] [PubMed]
  45. Scandolo, S.; Giannozzi, P.; Cavazzoni, C.; de Gironcoli, S.; Pasquarello, A.; Baroni, S. First-principles codes for computational crystallography in the Quantum-ESPRESSO package. Z. Krist. Cryst. Mater. 2005, 220, 574–579. [Google Scholar] [CrossRef]
  46. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef]
  47. Vanderbilt, D. Soft self-consistent pseudopotentials in a generalized eigenvalue formalism. Phys. Rev. B 1990, 41, 7892–7895. [Google Scholar] [CrossRef]
  48. Monkhorst, H.J.; Pack, J.D. Special points for Brillouin-zone integrations. Phys. Rev. B 1976, 13, 5188–5192. [Google Scholar] [CrossRef]
  49. Camellone, M.F.; Kowalski, P.M.; Marx, D. Ideal, defective, and gold-promoted rutile TiO2(110) surfaces interacting with CO, H2,and H2O: Structures, energies, thermodynamics, and dynamics from PBE + U. Phys. Rev. B 2011, 84, 035413. [Google Scholar] [CrossRef]
  50. Finazzi, E.; Di Valentin, C.; Pacchioni, G.; Selloni, A. Excess electron states in reduced bulk anatase TiO2: Comparison of standard GGA, GGA + U, and hybrid DFT calculations. J. Chem. Phys. 2008, 129, 154113. [Google Scholar] [CrossRef]
  51. Morgan, B.J.; Watson, G.W. A DFT + U description of oxygen vacancies at the TiO2 rutile (110) surface. Surf. Sci. 2007, 601, 5034–5041. [Google Scholar] [CrossRef]
  52. Di Valentin, C.; Pacchioni, G.; Selloni, A. Electronic Structure of Defect States in Hydroxylated and Reduced Rutile TiO2(110) Surfaces. Phys. Rev. Lett. 2006, 97, 166803. [Google Scholar] [CrossRef] [PubMed]
  53. Li, C.-G.; Shen, Z.-G.; Hu, Y.-F.; Tang, Y.-N.; Chen, W.-G.; Ren, B.-Z. Insights into the structures and electronic properties of Cu n + 1 μ and CunSμ (n = 1–12; μ = 0; ±1) clusters. Sci. Rep. 2017, 7, 1345. [Google Scholar] [CrossRef] [PubMed]
  54. Deák, P.; Kullgren, J.; Frauenheim, T. Polarons and oxygen vacancies at the surface of anatase TiO2. Phys. Status Solidi (RRL) Rapid Res. Lett. 2014, 8, 583–586. [Google Scholar] [CrossRef]
  55. Shibuya, T.; Yasuoka, K.; Mirbt, S.; Sanyal, B. Bipolaron Formation Induced by Oxygen Vacancy at Rutile TiO2 (110) Surfaces. J. Phys. Chem. C 2014, 118, 9429–9435. [Google Scholar] [CrossRef]
  56. Deák, P.; Aradi, B.; Frauenheim, T. Polaronic effects in TiO2 calculated by the HSE06 hybrid functional: Dopant passivation by carrier self-trapping. Phys. Rev. B 2011, 83, 155207. [Google Scholar] [CrossRef]
  57. Janotti, A.; Varley, J.B.; Rinke, P.; Umezawa, N.; Kresse, G.; Van de Walle, C.G. Hybrid functional studies of the oxygen vacancy in TiO2. Phys. Rev. B 2010, 81, 085212. [Google Scholar] [CrossRef]
  58. Kowalski, P.M.; Camellone, M.F.; Nair, N.N.; Meyer, B.; Marx, D. Charge localization dynamics induced by oxygen vacancies on the TiO2 (110) surface. Phys. Rev. Lett. 2010, 105, 146405. [Google Scholar] [CrossRef]
  59. Deskins, N.A.; Rousseau, R.; Dupuis, M. Localized Electronic States from Surface Hydroxyls and Polarons in TiO2 (110). J. Phys. Chem. C 2009, 113, 14583–14586. [Google Scholar] [CrossRef]
  60. Finazzi, E.; Di Valentin, C.; Pacchioni, G. Nature of Ti Interstitials in Reduced Bulk Anatase and Rutile TiO2. J. Phys. Chem. C 2009, 113, 3382–3385. [Google Scholar] [CrossRef]
  61. Li, S.; Jena, P. Origin of the anatase to rutile conversion of metal-doped TiO2. Phys. Rev. B 2009, 79, 201204. [Google Scholar] [CrossRef]
  62. Pan, H.; Gu, B.; Zhang, Z. Phase-Dependent Photocatalytic Ability of TiO2: A First-Principles Study. J. Chem. Theory Comput. 2009, 5, 3074–3078. [Google Scholar] [CrossRef] [PubMed]
  63. Deskins, N.A.; Dupuis, M. Electron transport via polaron hopping in bulk TiO2: A density functional theory characterization. Phys. Rev. B 2007, 75, 195212. [Google Scholar] [CrossRef]
  64. Sun, J.; Ruzsinszky, A.; Perdew, J.P. Strongly Constrained and Appropriately Normed Semilocal Density Functional. Phys. Rev. Lett. 2015, 115, 036402. [Google Scholar] [CrossRef] [PubMed]
  65. Tameh, M.S.; Dearden, A.K.; Huang, C. Accuracy of density functional theory for predicting kinetics of methanol synthesis from CO and CO2 hydrogenation on copper. J. Phys. Chem. C 2018, 122, 17942–17953. [Google Scholar] [CrossRef]
  66. Long, O.Y.; Gautam, G.S.; Carter, E.A. Evaluating optimal U for 3d transition-metal oxides within the SCAN + U framework. Phys. Rev. Mater. 2020, 4, 045401. [Google Scholar] [CrossRef]
  67. Lammers, U.; Borstel, G. Electronic and atomic structure of copper clusters. Phys. Rev. B 1994, 49, 17360–17377. [Google Scholar] [CrossRef]
  68. Bersani, M.; Gupta, K.; Mishra, A.K.; Lanza, R.; Taylor, S.F.; Islam, H.U.; Hollingsworth, N.; Hardacre, C.; de Leeuw, N.; Darr, J. Combined EXAFS, XRD, DRIFTS, and DFT Study of Nano Copper-Based Catalysts for CO2 Hydrogenation. ACS Catal. 2016, 6, 5823–5833. [Google Scholar] [CrossRef]
  69. Mishra, A.K.; Roldan, A.; de Leeuw, N.H. CuO Surfaces and CO2 Activation: A Dispersion-Corrected DFT + U Study. J. Phys. Chem. C 2016, 120, 2198–2214. [Google Scholar] [CrossRef]
  70. Song, Y.Y.; Wang, G.C. A DFT Study and Microkinetic Simulation of Propylene Partial Oxidation on CuO (111) and CuO (100) Surfaces. J. Phys. Chem. C 2016, 120, 27430–27442. [Google Scholar] [CrossRef]
  71. Scanlon, D.O.; Walsh, A.; Morgan, B.J.; Watson, G.W.; Payne, D.J.; Egdell, R.G. Effect of Cr substitution on the electronic structure of CuAl1−xCrxO2. Phys. Rev. B 2009, 79, 035101. [Google Scholar] [CrossRef]
  72. Moreira, I.D.P.; Rivero, P.; Illas, F. Electronic structure of HgBa2Can−1CunO2n+2 (n = 1, 2, 3) superconductor parent compounds from periodic hybrid density functional theory. J. Chem. Phys. 2011, 134, 074709. [Google Scholar] [CrossRef]
Figure 1. Variation of angle and interatomic distance for Cu 3 copper clusters.
Figure 1. Variation of angle and interatomic distance for Cu 3 copper clusters.
Condensedmatter 08 00081 g001
Figure 2. Energy as a function of the θ angle for the Cu 3 cluster for optimized values of the interatomic distance d. The Hubbard parameter is set to U = 5 eV.
Figure 2. Energy as a function of the θ angle for the Cu 3 cluster for optimized values of the interatomic distance d. The Hubbard parameter is set to U = 5 eV.
Condensedmatter 08 00081 g002
Figure 3. (a) Binding energies of the GS isomers vs. cluster size for the different levels of theory in this work. (b) Previous theoretical and experimental results for comparison, a Ref. [23], b Ref. [21], c Ref. [36], d Ref. [18], e Ref. [22], f Ref. [20] and g Ref. [17]. The U values are in eV.
Figure 3. (a) Binding energies of the GS isomers vs. cluster size for the different levels of theory in this work. (b) Previous theoretical and experimental results for comparison, a Ref. [23], b Ref. [21], c Ref. [36], d Ref. [18], e Ref. [22], f Ref. [20] and g Ref. [17]. The U values are in eV.
Condensedmatter 08 00081 g003
Figure 4. Density of states of phonons for Cu N copper clusters. U = 8 eV was used. Each letter i in the figure corresponds to the letter i of each cluster Cu N (i).
Figure 4. Density of states of phonons for Cu N copper clusters. U = 8 eV was used. Each letter i in the figure corresponds to the letter i of each cluster Cu N (i).
Condensedmatter 08 00081 g004
Figure 5. Density of States (DOS) for Cu N copper clusters. U = 5 eV was used. The Fermi energy was placed at 0 eV and corresponds to the dotted vertical line.
Figure 5. Density of States (DOS) for Cu N copper clusters. U = 5 eV was used. The Fermi energy was placed at 0 eV and corresponds to the dotted vertical line.
Condensedmatter 08 00081 g005
Table 1. Cu N clusters that were optimized in this work using atomic relaxations.
Table 1. Cu N clusters that were optimized in this work using atomic relaxations.
Condensedmatter 08 00081 i001Condensedmatter 08 00081 i002Condensedmatter 08 00081 i003Condensedmatter 08 00081 i004Condensedmatter 08 00081 i005Condensedmatter 08 00081 i006
GS Cu 3 (A)Cu 3 (B)GS Cu 4 (C)Cu 4 (D)Cu 4 (E)Cu 4 (F)
Condensedmatter 08 00081 i007Condensedmatter 08 00081 i008Condensedmatter 08 00081 i009Condensedmatter 08 00081 i010Condensedmatter 08 00081 i011Condensedmatter 08 00081 i012
GS Cu 5 (G)Cu 5 (H)Cu 5 (I)Cu 5 (J)Cu 5 (K)GS Cu 6 (L)
Condensedmatter 08 00081 i013Condensedmatter 08 00081 i014Condensedmatter 08 00081 i015Condensedmatter 08 00081 i016Condensedmatter 08 00081 i017Condensedmatter 08 00081 i018
Cu 6 (M)Cu 6 (N)Cu 6 (O)Cu 6 (P)Cu 6 (Q)Cu 6 (R)
GS: Ground state.
Table 2. Binding energies ( B E ) of Cu N (N = 3–4 atoms) isomers and their energy differences per atom ( Δ B E ) with respect to the GS as a function of the U parameter.
Table 2. Binding energies ( B E ) of Cu N (N = 3–4 atoms) isomers and their energy differences per atom ( Δ B E ) with respect to the GS as a function of the U parameter.
UCondensedmatter 08 00081 i019Condensedmatter 08 00081 i020Condensedmatter 08 00081 i021Condensedmatter 08 00081 i022Condensedmatter 08 00081 i023Condensedmatter 08 00081 i024
GS Cu 3 (A)Cu 3 (B)GS Cu 4 (C)Cu 4 (D)Cu 4 (E)Cu 4 (F)
B E Δ B E B E Δ B E B E Δ B E B E Δ B E B E Δ B E B E Δ B E
(eV)(eV)(meV)(eV)(meV)(eV)(meV)(eV)(meV)(eV)(meV)(eV)(meV)
01.36101.2541071.71801.622961.5441741.502217
11.27801.1621161.63201.536961.4481841.407225
21.20901.110991.55901.465941.3691901.327232
31.18401.101831.53001.438921.3361941.293237
41.16201.087751.50401.414901.3081971.262242
51.14301.074681.48201.394881.2831991.236246
61.12501.063621.46101.376851.2612001.212249
71.11001.053571.44301.359831.2422001.190252
81.09601.044511.42601.345811.2252001.171255
91.08301.035471.41001.331791.2092011.153257
HSE0.9950NCNC1.32201.234881.141182NCNC
DFT + PBE [21]1.433 I T 0- -- -1.85101.75596- -- -- -- -
DFT + PW [36]1.129 I T 0- -- -1.5050- -- -1.3291761.269236
DFT + BLYP [18]1.030 E T 00.962681.3070- -- -- -- -1.077230
Tight binding [19]1.660 E T 0- -- -- -- -- -- -- -- -1.93 G S 0
DFT + B3LYP [20]1.001 I T 0- -- -1.3000- -- -- -- -- -- -
DFT + PW91 [22]1.164 I T 0- -- -1.51501.42986- -- -- -- -
Monte Carlo, T = 300 K [17]1.1370- -- -- -- -- -- -- -- -1.460 G S 0
Exp. [23]1.0630- -- -1.4780- -- -- -- -- -- -
NC: The isomer does not converge. - -: Not reported. G S : Ground state. I T : isosceles triangle. E T : Equilateral triangle.
Table 3. Binding energies ( B E ) of Cu N (N = 5–6) isomers and their energy differences per atom ( Δ B E ) with respect to the GS as a function of the U parameter.
Table 3. Binding energies ( B E ) of Cu N (N = 5–6) isomers and their energy differences per atom ( Δ B E ) with respect to the GS as a function of the U parameter.
UCondensedmatter 08 00081 i025Condensedmatter 08 00081 i026Condensedmatter 08 00081 i027Condensedmatter 08 00081 i028Condensedmatter 08 00081 i029Condensedmatter 08 00081 i030
GS Cu 5 (G)Cu 5 (H)Cu 5 (I)Cu 5 (J)Cu 5 (K)GS Cu 6 (L)
B E Δ B E B E Δ B E B E Δ B E B E Δ B E B E Δ B E B E Δ B E
(eV)(eV)(meV)(eV)(meV)(eV)(meV)(eV)(meV)(eV)(meV)(eV)(meV)
01.85901.816431.813461.779801.7511082.0510
11.76901.718501.722471.678901.6641051.9650
21.69301.637571.647471.595991.5911021.8910
31.66201.600631.615471.5571051.563991.8590
41.63501.567681.588471.5241111.538971.8310
51.61101.538731.564471.4951151.516951.8060
61.58801.511771.542461.4701191.496931.7830
71.56901.488811.523461.4461221.478911.7650
81.55001.466851.505461.4261251.461891.7460
91.53301.446881.488451.4071271.446871.7260
HSE1.44901.382671.39951NCNC1.356931.6330
DFT + PBE [21]1.96301.91350- -- -- -- -- -- -2.1780
DFT + PW [36]1.63401.58054- -- -- -- -- -- -1.8340
DFT + BLYP [18]1.42801.34286- -- -1.318110- -- -1.5890
Tight binding [19]- -- -- -- -- -- -- -- -- -- -- -- -
DFT + B3LYP [20]1.2730- -- -- -- -- -- -- -- -1.5940
DFT + PW91 [22]1.64301.58558- -- -- -- -- -- -1.8340
Monte Carlo, T = 300 K [17]- -- -1.646 G S 0- -- -- -- -- -- -- -- -
Exp. [23]1.5520- -- -- -- -- -- -- -- -1.7200
NC: The isomer does not converge. - -: Not reported. G S : Ground state.
Table 4. Binding energies ( B E ) of Cu 6 isomers and their energy differences per atom ( Δ B E ) with respect to the GS as a function of the U parameter.
Table 4. Binding energies ( B E ) of Cu 6 isomers and their energy differences per atom ( Δ B E ) with respect to the GS as a function of the U parameter.
UCondensedmatter 08 00081 i031Condensedmatter 08 00081 i032Condensedmatter 08 00081 i033Condensedmatter 08 00081 i034Condensedmatter 08 00081 i035Condensedmatter 08 00081 i036
Cu 6 (M)Cu 6 (N)Cu 6 (O)Cu 6 (P)Cu 6 (Q)Cu 6 (R)
B E Δ B E B E Δ B E B E Δ B E B E Δ B E B E Δ B E B E Δ B E
(eV)(eV)(meV)(eV)(meV)(eV)(meV)(eV)(meV)(eV)(meV)(eV)(meV)
02.033192.014372.010421.973781.9051461.873179
11.949151.917471.923421.866991.8031621.777188
21.876151.836551.849411.7761151.7181731.698193
31.845141.798611.819401.7321271.6781811.663196
41.817141.765661.793391.6931381.6441871.633198
51.792141.735711.768381.6591471.6141921.607199
61.770141.708751.746371.6291551.5871971.583200
71.751141.683821.726381.6011631.5632021.566203
81.732141.661851.708381.5771691.5412051.542204
91.71411NCNC1.691351.5541721.5212051.524202
HSE1.617161.573591.59339NCNC1.4601721.447185
DFT + PBE [21]2.153242.136422.12850- -- -- -- -- -- -
DFT + PW [36]1.797371.75777- -- -1.6851491.637197- -- -
DFT + BLYP [18]1.548411.49792- -- -1.456133- -- -- -- -
Tight binding [19]- -- -- -- -- -- -2.200 G S 0- -- -- -- -
DFT + B3LYP [20]- -- -- -- -- -- -- -- -- -- -- -- -
DFT + PW91 [22]1.81222- -- -1.78549- -- -- -- -- -- -
Monte Carlo, T = 300 K [17]- -- -- -- -- -- -1.835 G S 0- -- -- -- -
NC: The isomer does not converge. - -: Not reported. G S : Ground state.
Table 5. Binding energy ordering of Cu N clusters for U = 0 .
Table 5. Binding energy ordering of Cu N clusters for U = 0 .
Cu 3 B E (A)> B E (B)
Cu 4 B E (C)> B E (D)> B E (E)> B E (F)
Cu 5 B E (G)> B E (H)> B E (I)> B E (J)> B E (K)
Cu 6 B E (L)> B E (M)> B E (N)> B E (O)> B E (P)> B E (Q)> B E (R)
Table 6. Binding energy ordering of Cu N clusters for U = 8 eV.
Table 6. Binding energy ordering of Cu N clusters for U = 8 eV.
Cu 3 B E (A)> B E (B)
Cu 4 B E (C)> B E (D)> B E (E)> B E (F)
Cu 5 B E (G)> B E (I)> B E (K)> B E (H)> B E (J)
Cu 6 B E (L)> B E (M)> B E (O)> B E (N)> B E (P)> B E (R)> B E (Q)
Table 7. Values of the bond lengths ( B L ), magnetization ( μ B ), polarization (P), and intragaps within the valence band ( Δ g ) of the copper clusters ( C u N , N = 3 5 ) for U = 0, 5, and 8 eV.
Table 7. Values of the bond lengths ( B L ), magnetization ( μ B ), polarization (P), and intragaps within the valence band ( Δ g ) of the copper clusters ( C u N , N = 3 5 ) for U = 0, 5, and 8 eV.
Condensedmatter 08 00081 i037Condensedmatter 08 00081 i038Condensedmatter 08 00081 i039Condensedmatter 08 00081 i040Condensedmatter 08 00081 i041
Cu 3 (A)Cu 4 (C)Cu 4 (D)Cu 5 (G)Cu 5 (H)
(1–2) = 2.333(1–2) = 2.368(1–2) = 2.241(1–2) = 2.376(1–2) = 2.414
(1–3) = 2.333(1–3) = 2.368(2–3) = 2.374(1–3) = 2.349(1–3) = 2.414
(2–3) = 2.333(2–3) = 2.272(2–4) = 2.374(1–4) = 2.386(1–4) = 2.414
(2–4) = 2.368(3–4) = 2.259(2–4) = 2.386(2–3) = 2.380
B L (Å) (3–4) = 2.368 (2–5) = 2.349(2–4) = 2.380
U = 0 (3–4) = 2.324(2–5) = 2.414
(4–5) = 2.324(3–4) = 2.380
(3–5) = 2.414
(4–5) = 2.414
(1–2) = 2.350(1–2) = 2.391(1–2) = 2.256(1–2) = 2.403(1–2) = 2.435
(1–3) = 2.350(1–3) = 2.391(2–3) = 2.409(1–3) = 2.361(1–3) = 2.435
(2–3) = 2.350(2–3) = 2.260(2–4) = 2.409(1–4) = 2.392(1–4) = 2.435
(2–4) = 2.391(3–4) = 2.257(2–4) = 2.392(2–3) = 2.384
B L (Å) (3–4) = 2.391 (2–5) = 2.361(2–4) = 2.383
U = 5 (3–4) = 2.347(2–5) = 2.435
(4–5) = 2.347(3–4) = 2.383
(3–5) = 2.435
(4–5) = 2.435
(1–2) = 2.360(1–2) = 2.402(1–2) = 2.267(1–2) = 2.416(1–2) = 2.447
(1–3) = 2.360(1–3) = 2.402(2–3) = 2.427(1–3) = 2.369(1–3) = 2.447
(2–3) = 2.360(2–3) = 2.261(2–4) = 2.427(1–4) = 2.399(1–4) = 2.447
(2–4) = 2.402(3–4) = 2.261(2–4) = 2.399(2–3) = 2.389
B L (Å) (3–4) = 2.402 (2–5) = 2.369(2–4) = 2.389
U = 8 (3–4) = 2.358(2–5) = 2.447
(4–5) = 2.358(3–4) = 2.389
(3–5) = 2.447
(4–5) = 2.447
μ B (a.u.), U = 0 1.060.000.000.931.13
μ B (a.u.), U = 5 1.070.000.001.001.13
μ B (a.u.), U = 8 1.070.000.001.011.13
(1) = 0.351(1) = 0.000(1) = 0.000(1) = 0.229(1) = 0.131
(2) = 0.352(2) = 0.000(2) = 0.000(2) = 0.229(2) = 0.286
P (a.u.), U = 0 (3) = 0.352(3) = 0.000(3) = 0.000(3) = 0.128(3) = 0.286
(4) = 0.000(4) = 0.000(4) = 0.212(4) = 0.286
(5) = 0.128(5) = 0.131
(1) = 0.352(1) = 0.000(1) = 0.000(1) = 0.246(1) = 0.129
(2) = 0.352(2) = 0.000(2) = 0.000(2) = 0.246(2) = 0.286
P (a.u.), U = 5 (3) = 0.352(3) = 0.000(3) = 0.000(3) = 0.136(3) = 0.286
(4) = 0.000(4) = 0.000(4) = 0.226(4) = 0.286
(5) = 0.136(5) = 0.129
(1) = 0.353(1) = 0.000(1) = 0.000(1) = 0.247(1) = 0.127
(2) = 0.353(2) = 0.000(2) = 0.000(2) = 0.247(2) = 0.287
P (a.u.), U = 8 (3) = 0.352(3) = 0.000(3) = 0.000(3) = 0.137(3) = 0.287
(4) = 0.000(4) = 0.000(4) = 0.227(4) = 0.287
(5) = 0.137(5) = 0.127
Δ g (eV), U = 0 0.000.000.000.000.00
Δ g (eV), U = 5 1.600.900.600.141.00
Δ g (eV), U = 8 1.801.201.100.601.90
Table 8. Values of the bond lengths ( B L ) of the copper clusters (Cu 6 ) for U = 0, 5, and 8 eV.
Table 8. Values of the bond lengths ( B L ) of the copper clusters (Cu 6 ) for U = 0, 5, and 8 eV.
Condensedmatter 08 00081 i042Condensedmatter 08 00081 i043Condensedmatter 08 00081 i044Condensedmatter 08 00081 i045Condensedmatter 08 00081 i046Condensedmatter 08 00081 i047
Cu 6 (L)Cu 6 (M)Cu 6 (N)Cu 6 (O)Cu 6 (P)Cu 6 (R)
(1–2) = 2.320(1–2) = 2.417(1–2) = 2.370(1–2) = 2.376(1–2) = 2.413(1–2) = 2.344
(1–3) = 2.320(1–3) = 2.342(1–3) = 2.480(1–3) = 2.403(1–3) = 2.413(1–4) = 2.362
(2–3) = 2.411(1–4) = 2.417(1–4) = 2.458(1–4) = 2.312(1–4) = 2.413(1–5) = 2.429
(2–4) = 2.322(2–3) = 2.342(1–6) = 2.458(2–4) = 2.305(1–5) = 2.413(2–3) = 2.324
B L (Å)(2–5) = 2.411(2–5) = 2.417(2–4) = 2.458(3–4) = 2.414(2–3) = 2.413(2–5) = 2.375
U = 0 (3–5) = 2.411(3–4) = 2.342(2–5) = 2.480(3–5) = 2.403(2–5) = 2.413(2–6) = 2.429
(3–6) = 2.322(3–5) = 2.342(2–6) = 2.458(4–5) = 2.312(2–6) = 2.413(3–6) = 2.362
(4–5) = 2.321(3–6) = 2.342(3–4) = 2.378(4–6) = 2.305(3–4) = 2.413(4–5) = 2.324
(5–6) = 2.321(4–6) = 2.417(3–6) = 2.378(5–7) = 2.376(3–6) = 2.413(5–6) = 2.344
(5–6) = 2.417(4–5) = 2.378 (4–5) = 2.413
(4–6) = 2.320 (4–6) = 2.413
(5–6) = 2.378 (5–6) = 2.413
(1–2) = 2.349(1–2) = 2.457(1–2) = 2.375(1–2) = 2.410(1–2) = 2.427(1–2) = 2.383
(1–3) = 2.349(1–3) = 2.338(1–3) = 2.523(1–3) = 2.438(1–3) = 2.427(1–4) = 2.370
(2–3) = 2.420(1–4) = 2.457(1–4) = 2.483(1–4) = 2.315(1–4) = 2.427(1–5) = 2.394
(2–4) = 2.349(2–3) = 2.338(1–6) = 2.484(2–4) = 2.310(1–5) = 2.427(2–3) = 2.366
B L (Å)(2–5) = 2.420(2–5) = 2.457(2–4) = 2.483(3–4) = 2.403(2–3) = 2.428(2–5) = 2.430
U = 5 (3–5) = 2.420(3–4) = 2.338(2–5) = 2.523(3–5) = 2.438(2–5) = 2.428(2–6) = 2.394
(3–6) = 2.349(3–5) = 2.338(2–6) = 2.484(4–5) = 2.315(2–6) = 2.428(3–6) = 2.370
(4–5) = 2.349(3–6) = 2.338(3–4) = 2.392(4–6) = 2.310(3–4) = 2.428(4–5) = 2.366
(5–6) = 2.349(4–6) = 2.457(3–6) = 2.392(5–7) = 2.410(3–6) = 2.428(5–6) = 2.283
(5–6) = 2.457(4–5) = 2.392 (4–5) = 2.428
(4–6) = 2.292 (4–6) = 2.428
(5–6) = 2.392 (5–6) = 2.428
(1–2) = 2.359(1–2) = 2.472(1–2) = 2.380(1–2) = 2.426(1–2) = 2.439(1–2) = 2.395
(1–3) = 2.359(1–3) = 2.342(1–3) = 2.545(1–3) = 2.458(1–3) = 2.439(1–4) = 2.379
(2–3) = 2.427(1–4) = 2.472(1–4) = 2.497(1–4) = 2.312(1–4) = 2.439(1–5) = 2.396
(2–4) = 2.359(2–3) = 2.342(1–6) = 2.499(2–4) = 2.313(1–5) = 2.439(2–3) = 2.381
B L (Å)(2–5) = 2.427(2–5) = 2.472(2–4) = 2.497(3–4) = 2.401(2–3) = 2.439(2–5) = 2.448
U = 8 (3–5) = 2.427(3–4) = 2.342(2–5) = 2.545(3–5) = 2.458(2–5) = 2.439(2–6) = 2.528
(3–6) = 2.359(3–5) = 2.342(2–6) = 2.499(4–5) = 2.319(2–6) = 2.439(3–6) = 2.379
(4–5) = 2.359(3–6) = 2.342(3–4) = 2.400(4–6) = 2.313(3–4) = 2.439(4–5) = 2.381
(5–6) = 2.359(4–6) = 2.472(3–6) = 2.400(5–7) = 2.426(3–6) = 2.439(5–6) = 2.395
(5–6) = 2.472(4–5) = 2.400 (4–5) = 2.439
(4–6) = 2.289 (4–6) = 2.439
(5–6) = 2.400 (5–6) = 2.439
Table 9. Values of the magnetization ( μ B ), polarization (P), and intragaps within the valence band ( Δ g ) of the copper clusters ( C u N , N = 3–5) for U = 0, 5, and 8 eV.
Table 9. Values of the magnetization ( μ B ), polarization (P), and intragaps within the valence band ( Δ g ) of the copper clusters ( C u N , N = 3–5) for U = 0, 5, and 8 eV.
Condensedmatter 08 00081 i048Condensedmatter 08 00081 i049Condensedmatter 08 00081 i050Condensedmatter 08 00081 i051Condensedmatter 08 00081 i052Condensedmatter 08 00081 i053
Cu 6 (L)Cu 6 (M)Cu 6 (N)Cu 6 (O)Cu 6 (P)Cu 6 (R)
μ B (a.u.), U = 0 0.000.000.000.002.121.42
μ B (a.u.), U = 5 0.000.000.000.002.211.91
μ B (a.u.), U = 8 0.000.000.000.002.211.95
(1) = 0.000(1) = 0.000(1) = 0.000(1) = 0.000(1) = 0.351(1) = 0.228
(2) = 0.000(2) = 0.000(2) = 0.000(2) = 0.000(2) = 0.351(2) = 0.235
P (a.u.), U = 0 (3) = 0.000(3) = 0.000(3) = 0.000(3) = 0.000(3) = 0.351(3) = 0.239
(4) = 0.000(4) = 0.000(4) = 0.000(4) = 0.000(4) = 0.351(4) = 0.239
(5) = 0.000(5) = 0.000(5) = 0.000(5) = 0.000(5) = 0.351(5) = 0.235
(6) = 0.000(6) = 0.000(6) = 0.000(6) = 0.000(6) = 0.351(6) = 0.228
(1) = 0.000(1) = 0.000(1) = 0.000(1) = 0.000(1) = 0.365(1) = 0.304
(2) = 0.000(2) = 0.000(2) = 0.000(2) = 0.000(2) = 0.365(2) = 0.321
P (a.u.), U = 5 (3) = 0.000(3) = 0.000(3) = 0.000(3) = 0.000(3) = 0.365(3) = 0.320
(4) = 0.000(4) = 0.000(4) = 0.000(4) = 0.000(4) = 0.365(4) = 0.320
(5) = 0.000(5) = 0.000(5) = 0.000(5) = 0.000(5) = 0.365(5) = 0.321
(6) = 0.000(6) = 0.000(6) = 0.000(6) = 0.000(6) = 0.365(6) = 0.304
(1) = 0.000(1) = 0.000(1) = 0.000(1) = 0.000(1) = 0.365(1) = 0.308
(2) = 0.000(2) = 0.000(2) = 0.000(2) = 0.000(2) = 0.365(2) = 0.330
P (a.u.), U = 8 (3) = 0.000(3) = 0.000(3) = 0.000(3) = 0.000(3) = 0.365(3) = 0.326
(4) = 0.000(4) = 0.000(4) = 0.000(4) = 0.000(4) = 0.365(4) = 0.326
(5) = 0.000(5) = 0.000(5) = 0.000(5) = 0.000(5) = 0.365(5) = 0.330
(6) = 0.000(6) = 0.000(6) = 0.000(6) = 0.000(6) = 0.365(6) = 0.308
Δ g (eV), U = 0 0.000.000.000.600.000.00
Δ g (eV), U = 5 1.200.600.601.001.100.50
Δ g (eV), U = 8 0.901.401.600.702.200.70
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Alcalá-Varilla, L.A.; Ponnefz-Durango, R.E.; Seriani, N.; Araujo-Lopez, E.; Montoya, J.A. A DFT + U Study on the Stability of Small CuN Clusters (N = 3–6 Atoms): Calculation of Phonon Frequencies. Condens. Matter 2023, 8, 81. https://doi.org/10.3390/condmat8030081

AMA Style

Alcalá-Varilla LA, Ponnefz-Durango RE, Seriani N, Araujo-Lopez E, Montoya JA. A DFT + U Study on the Stability of Small CuN Clusters (N = 3–6 Atoms): Calculation of Phonon Frequencies. Condensed Matter. 2023; 8(3):81. https://doi.org/10.3390/condmat8030081

Chicago/Turabian Style

Alcalá-Varilla, Luis A., Rafael E. Ponnefz-Durango, Nicola Seriani, Eduard Araujo-Lopez, and Javier A. Montoya. 2023. "A DFT + U Study on the Stability of Small CuN Clusters (N = 3–6 Atoms): Calculation of Phonon Frequencies" Condensed Matter 8, no. 3: 81. https://doi.org/10.3390/condmat8030081

Article Metrics

Back to TopTop