Structural stability and bonding nature of Li–Sn–carbon nanocomposites as Li-ion battery anodes: first principles approach

T. K. Bijoya, J. Karthikeyanb and P. Murugan*ab
aCSIR-Network Institute of Solar Energy (CSIR-NISE), CSIR-Central Electrochemical Research Institute, Karaikudi-630 006, Tamil Nadu, India. E-mail: murugan@cecri.res.in
bFunctional Materials Division, CSIR-Central Electrochemical Research Institute, Karaikudi-630 006, Tamil Nadu, India

Received 25th September 2014 , Accepted 13th November 2014

First published on 13th November 2014


Abstract

The atomic structural stability and electronic properties of LinSn4–carbon nanotube (CNT) and LinSn4–graphene nanocomposites were studied by first principles calculations. Results on isolated LinSn4 clusters, with n = 0–10, revealed that the tetrahedron shaped Li4Sn4 Zintl cluster is the most stable owing to it having high symmetry as well as a largest highest occupied molecular orbital–lowest unoccupied molecular orbital (HOMO–LUMO) gap. This LinSn4 cluster weakly interacted with CNT as well as graphene for n ≤ 4, whereas a strong cation–π interaction is observed between them for n > 4 which significantly reduces the Li clustering. The interaction between the Sn cluster and CNT or graphene is mediated only through Li ions whose absence destabilizes the Sn–C composite. These results were further confirmed by electronic density of states and band structure calculations. In addition, our calculations on hexagonal assembly of LinSn4–CNT imply that the volume change is minimal during lithiation and the average intercalation potential is estimated to be a maximum of ≈0.5 V, which shows its good anodic character.


1. Introduction

Usage of carbonaceous materials in energy storage applications has increased tremendously since the discovery of several carbon nanostructures, such as carbon nanotubes (CNT),1 C60 bucky ball,2 graphene,3 and carbon nanofibers.4 Among various energy storage devices, lithium ion batteries (LIB) are widely used in many portable electronic devices5 due to their high energy density and light weight, which make them of interest to scientific as well as industrial researchers. Several attempts have been made to design good materials for the components of lithium ion batteries, which are essentially the electrodes, the electrolyte, and the separator. In particular, materials for anodes have been focused in recent years because the specific capacity could possibly be raised to 4200 mA h g−1 (theoretical capacity of Si).6,7 Although the theoretical capacity of silicon has been found to be high, it has a severe problem of poor inter particle conductivity due to fractures in the particle's structure, which can be minimized by controlling particle size.8 Graphite is a common material used for anodes due to its low cost and long battery life.9 However, six carbon atoms together can capture only one Li ion to form stable LiC6 stoichiometry, which results in a low theoretical specific capacity of 372 mA h g−1,10 while graphene with a few layers carries a higher capacity of 700 mA h g−1.11 Compared to this, defect-free single walled CNTs (SWCNTs) have a specific capacity of only 400–460 mA h g−1, but this reaches 1000 mA h g−1 for side-wall defective CNTs12 and this enhancement was reported to be due to the presence of high porosity and one-dimensionality. However, all these carbon nanostructures have several problems, such as the capacity fading during battery cycling and limitations in the capacity value. Hence, there are many attempts to replace the carbon with some other group IV elements such as Si, Sn, and Pb.13–16 It is worth noting that these elements are capable of forming a Zintl phase compound with alkali metal atoms which act as the cation and Sin, Snn or Pbn clusters as the anion.17,18 Compared to carbon, the Sn atom has the ability to bind with more Li ions, which results in the formation of a Li4.4Sn compound with a high theoretical capacity of 994 mA h g−1. However, the drawback in Sn compounds is the problem of a colossal volume change during the intercalation process, which leads to poor battery life.19

Concerning the battery life, it is important to find a material that exhibits good capacity with minimal volume change during cycling. Composites of nanostructured Sn with multiwalled CNT20 and carbon nanofibres21 have been synthesized with a greater specific capacity of 889 mA h g−1 with minimal volume change. Following these works, several reports22–32 were published in which the Sn–C based composites were shown to be potential candidates for LIB anodes. Although theoretical understanding of mono-atomic thinned Sn nanowire with SWCNT has been reported,31 the structural stability of the Li–Sn alloy and its interactions with CNT have not been studied well. Our calculations shed light on the atomic structural stability and electronic properties of the Li–Sn–CNT composite within density functional theory formalism. We initially studied the growth of Li–Sn alloyed clusters and then stable clusters were identified and inserted into various sized CNTs in order to understand the interaction between them. We also carried out the deposition of those clusters onto graphene. Our results on LinSn4 clusters with n = 0–10 reveal that the Li4Sn4 cluster has ultra-high stability with the largest highest occupied molecular orbital–lowest unoccupied molecular orbital (HOMO–LUMO) gap as it reaches the Zintl's composition and hence, it interacts weakly with CNT as well as graphene, whereas the clusters with more Li ions (n > 4) bind strongly as a result of significant charge transfer from cluster to CNT, through cation–π interactions. These results were further confirmed by electronic density of states and band structure calculations. In addition, we performed the calculations on the assembly of Sn–Li cluster inserted CNTs revealing that the volume change is minimal during lithiation and the average intercalation voltage (AIV) was found to be a maximum of ≈0.5 V which shows its good anodic character.

2. Computational methods

First principles density functional calculations were performed to understand the structural and electronic properties of Li–Sn cluster inserted SWCNT as implemented in the Vienna Ab initio Simulation Package (VASP).33 In our calculations, all the atoms were described by projector augmented wave pseudopotential method34 and electron–electron interaction was correlated by generalized gradient approximations.35 For LinSn4 clusters, a cubic box with a = 15 Å is used to give enough vacuum space between the cluster and its periodic images, to avoid the interaction between them. In the case of isolated CNTs, sufficiently large a and b lattice constants were chosen so that CNT and its periodic images are at least 9 Å apart. Along the z-direction, four super cells were constructed with periodicity maintained. For sampling the Brillouin zone, Monkhorst–Pack 1 × 1 × 1, 1 × 1 × 4, and 2 × 2 × 4 k-meshes were used for optimizing the Li–Sn clusters, cluster inserted CNT, and its assembly, respectively. We also repeated the calculations for clusters with a denser k-mesh (3 × 3 × 3) to ensure its accuracy and negligible energy difference, with a maximum of 0.025 eV per atom observed when compared to Γ-point calculations. Further, a 1 × 1 × 50 k-mesh was employed to obtain the density of states (DOS) and band structure of all the systems. All the ions were relaxed self-consistently without considering any symmetry and the iterative relaxation process was repeated until the absolute force on each ion converged to the order of 0.01 eV Å−1. The convergence for energy was also set to be 10−5 eV throughout our calculations. We repeated all calculations with the above-mentioned criteria to find the magnetic solution of the system, which is favoured for clusters with odd number of electrons. Note that the magnetism in this cluster is completely quenched when it is inserted into SWCNT as significant charge transfer between these two systems is observed from our calculations.

3. Results and discussion

We initially considered the Sn4 cluster to understand the lithiation process on it by first principles calculations. Among various isomers, the Sn4 rhombus is found to be more stable by 5 meV per atom than the next most stable tetrahedron isomer, which is in accordance with earlier work.36 Further, for addition of Li ions on the Sn4 cluster, both rhombus and tetrahedron isomers (shown in Fig. 1) are considered as this small energy difference between these isomers will be compensated by the Sn–Li alloying energy. First, a single Li ion is introduced into various possible sites such as the terminal, edge, and face. Our calculations show that Li ion prefers to occupy the facial site (shown in Fig. 1) rather than the edge and terminal sites. The terminal addition is energetically least feasible by 0.22 eV as compared to facial addition; however, the energy of facial and edge Li ion added isomers differ by only 0.01 eV. Hence, a further three more Li ions are independently introduced one by one, on the faces and edges of the isomers. Interestingly, it is observed that the optimized geometry of both Li4Sn4 isomers converges into a tetrahedron structure. This cluster is commonly known as a Zintl cluster and it has an Sn44− anion37 which is surrounded by four Li ions acting as cations. By addition of further Li ions on the edge (the next feasible site) of this cluster, the geometry is transformed to a butterfly like structure38,39 for Li6Sn4 and becomes planar for Li8Sn4 (Fig. 1). The above discussions indicate that the addition of Li ions strengthens the Sn–Sn bonds by donating electrons to the Sn–Sn bonding orbitals until n = 4; beyond this, the geometry of the cluster is finally converted to a planar structure with larger Sn–Sn bonds, which is attributed by the filling of Sn–Sn anti-bonding orbitals (Fig. S1, ESI). In addition, clustering of Li ions is also observed for the n > 4 case (Table 1). The calculated Sn–Sn, Sn–Li, and Li–Li bond distances are reported in Table 1 and these values are quite consistent with the earlier work.38,39 Thus, our study reveals that the structural change in the pure Sn cluster is unavoidable during the Sn–Li alloying process. Such structural changes affect the battery cycle life.
image file: c4ra11187g-f1.tif
Fig. 1 The ball and stick models of optimized geometries of stable LinSn4 clusters are shown along with the Sn4 tetrahedron and rhombus isomers. The n value of each cluster is given in parentheses. Yellow and green balls represent Sn and Li atoms, respectively.
Table 1 The average Sn–Sn, Sn–Li, Li–Li, and C–Li bond distances are given for Li–Sn clusters inserted into CNT or deposited on graphene along with Ead and the Bader charge per Sn atom (QB). Values for isolated clusters are provided in parentheses
System cluster/CNT Sn–Sn Å Sn–Li Å Li–Li Å C–Li Å Ead eV QB e
Sn4/(8,8) 3.04(2.90) −2.37 4.00(4.00)
Li1Sn4/(8,8) 2.87(2.93) 3.08(2.95) 2.44 0.43 4.08(4.22)
Li2Sn4/(8,8) 2.87(2.89) 2.93(2.92) 2.42 0.41 4.40(4.50)
Li3Sn4/(8,8) 2.87(2.95) 3.00(2.86) 2.45 0.80 4.44(4.75)
Li4Sn4/(8,8) 3.04(3.06) 2.96(2.81) 2.58 0.50 4.75(5.00)
Li6Sn4/(8,8) 3.02(3.09) 2.95(2.78) 3.25(3.08) 2.5 3.02 4.92(5.50)
Li8Sn4/(9,9) 3.05(3.02) 2.90(2.79) 3.12(3.08) 2.50 2.44 5.41(6.00)
Sn4/Gr 3.05(2.90) −1.21 4.00(4.00)
Li2Sn4/Gr 2.85(2.89) 3.03(2.92) 2.53 0.91 4.44(4.50)
Li4Sn4/Gr 3.01(3.06) 2.91(2.81) 2.65 0.11 4.73(5.00)
Li6Sn4/Gr 3.04(3.09) 2.87(2.78) 3.23(3.08) 2.39 1.55 4.98(5.50)
Li8Sn4/Gr 3.05(3.02) 2.89(2.79) 2.96(3.08) 2.45 1.02 5.46(6.00)


The binding energies per atom (BE) of the LinSn4 clusters were calculated to understand their structural stability from:

 
image file: c4ra11187g-t1.tif(1)
where E(LinSn4) is the total energy of the LinSn4 cluster and E(Sn) and E(Li) are atomic energies of Sn and Li atoms, respectively. The calculated BE values are reported in Fig. 2. The results imply that the BE of LinSn4 clusters decreases with n. Note that all clusters with even n values were found to be more stable than the clusters with odd n due to the presence of unpaired electron in the HOMO of the latter clusters. Among all the clusters, the Li4Sn4 cluster shows the highest stability and largest HOMO–LUMO energy gap, which is consistent with earlier work.38 Such high stability in this cluster is due to the presence of Td symmetry and the completely filled bonding orbitals (see Fig. S1, ESI).


image file: c4ra11187g-f2.tif
Fig. 2 BE and HOMO–LUMO gap of LinSn4 clusters with n = 0–10.

To understand the interaction between the clusters and CNT, pristine Sn4, and LinSn4 (n = 0–8) clusters were chosen to insert into various sized (m,m) armchair CNTs, where m is the chiral index varying from 7 to 10. The adsorption energy (Ead) is calculated from:

 
Ead = E(LinSn4) + E(CNT) − E(LinSn4 + CNT) (2)
where, E(LinSn4), E(CNT), and E(LinSn4 + CNT) are total energies of the LinSn4 cluster, pristine (m,m) armchair CNT, and LinSn4 inserted CNT, respectively. The Ead values are reported in Table 1 and for pristine Sn4 cluster loaded CNTs are observed to be negative for all the CNTs (e.g. −2.37 eV for (8,8) CNT), whereas this value is positive for all lithiated clusters (see Fig. S2, ESI). This result clearly indicates that in the absence of Li ions, the Sn cluster will not bind strongly with CNT. Our calculations on the preferential radius for inserting clusters into CNT show that (8,8) CNT is apt for clusters with n = 0, 2, 4, and 6, while (9,9) CNT is apt for the Li8Sn4 cluster. Then, the calculated Ead values for Sn4, Li2Sn4, Li4Sn4, Li6Sn4, and Li8Sn4 clusters are −2.37, 0.41, 0.52, 3.02, and 2.44 eV, respectively. Thus, our results revealed that the Ead value is drastically increased for n > 4 (after the formation of the Zintl cluster). In addition, we have also studied the insertion of clusters into larger sized CNT, in which the LinSn4 cluster prefers to interact with one side of the CNT's wall. (see Fig. S3, ESI†). As the result of lesser interactions, the Ead values are lowered to 0.32 and 2.19 eV for n = 4 and 6, respectively. Thus, we concluded that the interactions between Sn clusters and carbon nanostructures are mediated only through Li ions, whose absence destabilizes the Sn–C composite. Of (10,10) CNT's wall (see Fig. S3, ESI). As the result of lesser interactions, the Ead values are lowered to 0.32.

To demonstrate the interaction of these clusters with other types of CNTs, we also inserted the Li8Sn4 cluster into (12,6) chiral and (15,0) zigzag CNTs which have a similar radius to the (9,9) armchair CNT (see Fig. S4, ESI). The calculated Ead values are 2.02 and 2.56 eV, respectively, which are less than that of the armchair CNT. The lowering in interaction is due to the semiconducting nature of the chiral and zigzag CNTs (see Fig. S5, ESI). Hence, it is concluded that the size and electronic properties of the cluster and CNTs play vital roles in the stability of Sn–Li–CNT composites. These results induced us to study the interaction between a planar honeycomb network (graphene) and the lithiated Sn4 clusters. The calculated Ead of LinSn4/Gr (n = 0, 2, 4, 6 and 8) systems (optimized structures are shown in Fig. S6 in the ESI) follow the same trend with smaller values (see Table 1), owing to the planar geometry of the graphene sheet. As stronger interactions are observed between the CNTs and clusters with n > 4, it is also expected that structural changes in clusters and/or CNTs will be observed. For instance, in the Li6Sn4/(8,8) CNT system, the butterfly like structure of cluster reverts back to tetrahedral as it transfers two excess electrons to the CNT. Note that among the Li–Sn clusters, the Li4Sn4 tetrahedron cluster is highly stable. On the other hand, for the Li8Sn4/(9,9) CNT case, the geometry of the cluster is not affected much as all four electrons (compared to the Li4Sn4 cluster) could not be accepted by the CNT. In addition, it was also observed that the Li clusterization decreased when LinSn4 with n > 4 clusters were inserted inside the CNT. Such a reduction in Li dimers/clusters will be beneficial for the electrode system.40

For example, the Li–Li bond distances in Li6Sn4 and Li8Sn4 clusters were observed to change from 3.08 Å to 3.25 Å and 3.08 Å to 3.12 Å, respectively, when these clusters were inserted into CNT (Table 1). Thus, our study shows that if the cluster has more Li ions, the shape of the cluster does not change much when it is inserted inside the CNT. As expected, the shape of the CNT also changes when the Li8Sn4 cluster is inserted into the (9,9) CNT which is transformed from a regular to an elliptic cylinder shape, with a major to minor axis ratio (a/b) of 1.07 and deformation (|ab|) of 0.82 Å. However, to understand the structural changes in the cluster loaded multiwalled CNTs, we also studied double walled CNT (DWCNT) with inter wall distance of ≈3.47 Å, loaded with a Li8Sn4 cluster (the details are given in Fig. S7, ESI). Our results show that the a/b ratio and |ab| of the inner wall are only 1.02 and 0.26 Å, respectively. Hence, we conclude that the outer walls resist the deformation of the inner wall during the cluster loading in the case of multiwalled CNT.

To understand the nature of the interaction qualitatively, the charge transfer between the cluster and the CNT was plotted for all cases and is shown in Fig. 3. For the lithiated clusters, a disc like isosurface was observed between the CNT and the LinSn4 cluster, which explains the charge transfer from the Sn-5p orbitals to the π orbitals of the sp2 hybridized carbon atoms of CNT via Li ions. A similar charge transfer diagram was observed in Fig. S5 (ESI) for Li ions deposited on the surface of graphene41 and it closely resembles the cation–π interaction which was elaborately discussed.42 Although no charge transfer between the bare Sn4 cluster and the CNT was noticed, the charge re-population in CNT as well as the cluster was observed (see Fig. 3). Note that though the adsorption energy was found to be small for the semiconducting CNT and graphene surface, an almost similar charge transfer mechanism was obtained. Overall, this significant charge transfer from Sn–Li clusters to CNT, through cation–π interaction, is attributed to the good structural stability of the Sn–Li–CNT composites.


image file: c4ra11187g-f3.tif
Fig. 3 Red and blue isosurfaces indicate excess and depletion of charge in (a) Sn4/(8,8) CNT, (b) Li2Sn4/(8,8) CNT, (c) Li4Sn4/(8,8) CNT, (d) Li6Sn4/(8,8) CNT, (e) Li8Sn4/(9,9) CNT, and (f) Li6Sn4/Gr systems. Charge transfer (δρ) is calculated from: δρ = ρcluster + ρCNTρcluster+CNT where ρcluster,ρCNT and ρcluster+CNT are charge densities of cluster, CNT, and cluster loaded CNT, respectively.

Further, to quantify this charge transfer between the LinSn4 cluster and the CNT, Bader analysis43 was performed to estimate the valence charge on each atom and is reported in Table 1. It shows that there is no charge transfer between the bare Sn4 cluster and the CNT as the valence charge on the Sn atoms remains the same before and after insertion into the CNT which is consistent with the charge transfer plotted in Fig. 3. In the case of Li4Sn4/(8,8) CNT, 1.0 e was transferred from Sn-5p orbitals to C-2p orbitals through Li ions, whereas for Li6Sn4 and Li8Sn4 clusters, 2.32 and 2.36 e were transferred, respectively (Table 1). Hence, the CNT is able to draw the charge from the LinSn4 cluster with a maximum limit of ≈2.3 e. Even though there is nearly equal charge transfer for the n = 6 and 8 clusters to CNT, we could observe more Li dimers for the latter case which reduces the Ead values. The calculated Bader charges for the LinSn4/Gr systems show that there is no charge transfer from the bare Sn4 cluster to the graphene substrate, whereas comparatively fewer electrons are drawn from other clusters due to fewer C–Li bonds resulting from the planar geometry of the substrate.

To explore the electronic properties of LinSn4/CNT systems, the density of states (DOS) for all systems are deduced and reported in Fig. 4. It is observed in all cases that the sp2 hybridized states of carbon atoms are distributed in the energy range from −12 eV to −9 eV, whereas the C-2p bands are spread over the entire range (−12 eV to −1 eV). It is also noticed that the states of Sn-5s (not shown in Fig. 4) and 5p states are pronounced as narrow peaks because the electrons in these states are well localized due to the ionic character of the Sn–Li bond. The DOS of bare Sn4/(8,8) CNT shows that the occupied states are shifted (as indicated by arrows) slightly towards Fermi energy (Ef) when compared to pristine CNT which again manifests the poor stability of the Sn–CNT composite in the absence of lithium. Conversely, the total occupied DOS of LinSn4–CNT systems with n = 2, 4, 6, and 8 are shifted away from Ef by 0.07, 0.20, 0.78 and 0.73 eV, respectively, which give an obvious reason for the good stability of the Sn–Li–CNT composites (Fig. 4).


image file: c4ra11187g-f4.tif
Fig. 4 Total DOS (black solid line) of (a) Sn4/(8,8) CNT, (b) Li2Sn4/(8,8) CNT, (c) Li4Sn4/(8,8) CNT, (d) Li6Sn4/(8,8) CNT, and (e) Li8Sn4/(9,9) CNT systems. C-2s, C-2p, and Sn-5p states are represented by red, blue, and green solid lines, respectively. The total DOS of bare CNT is given in the background and the direction of arrows depicts the shifting of energy levels, referenced to bare CNT.

image file: c4ra11187g-f5.tif
Fig. 5 Electronic band structure of (a) Sn4/(8,8) CNT, (b) Li2Sn4/(8,8) CNT, (c) Li4Sn4/(8,8) CNT, (d) Li6Sn4/(8,8) CNT, and (e) Li8Sn4/(9,9) CNT systems. Solid dots represent electron occupancies in the Sn-5p levels.

The band structure calculations were performed for cluster inserted CNT systems to understand the electron occupancy in the Sn-5p levels and the filling of the C-2p bands (Fig. 5). In the Sn4/(8,8) CNT system, two bands originated from C-2pz orbitals cross the Ef and form a Dirac cone like feature at Ef + 0.5 eV. The singly and triply degenerated localized Sn-5p states are completely occupied and observed at Ef −0.5 eV and Ef, respectively, whereas two unoccupied bonding Sn-5p states lie closer to Ef +0.5 eV. These empty states get filled when the bare Sn4 cluster is lithiated. For instance, five Sn-5p states are occupied for the Li2Sn4/(8,8) CNT system and the C-2p bands are further filled, but not completely, due to the limited charge transfer from the Li ions. Conversely, the Li4Sn4/(8,8) CNT and Li6Sn4/(8,8) CNT systems have the same number of occupied Sn-5p states, because two electrons in the antibonding states of the latter system are transferred to fill the C-2p bands completely which were partially filled in the former case. Because of this charge transfer, the Li6Sn4 cluster changes both the geometry and the electronic structure to similar to that of the Li4Sn4 Zintl cluster owing to having a large HOMO–LUMO gap. Note that the C-2p bands are completely filled in this case. Hence, the Li8Sn4 cluster is not able to transfer all four excess electrons (compared to Li4Sn4 cluster) to CNT's C-2p bands; instead only two electrons are transferred, similar to the Li6Sn4 case. As the result of this lower charge transfer, the cluster is not able to attain a stable state. Moreover, in this case, the Sn-5p level of the cluster is well separated which indicates that the cluster reduces its symmetry due to having a planar structure. Hence, we conclude that the CNT is not able to draw all the electrons from the LinSn4 with (n > 6) cluster that is electronically limited by the completely filled C-2p bands of CNT.

The results of the cluster filled CNTs motivated us to form a hexagonal assembly to study the Li-ion intercalation properties (see Fig. S7, ESI). Generally, Li ions prefer to occupy the voids of the assembly. Here, Li ions are added one by one in well separated sites to reduce the columbic repulsion between them. The average intercalation voltage (AIV)44 is calculated using the Nernst equation:

 
image file: c4ra11187g-t2.tif(3)
where E(Lix1@LinSn4) and E(Lix2@LinSn4) are the total energies of x1 and x2 Li ions intercalated in the LinSn4 cluster inserted into (m,m) CNT, respectively. E(Li) and F are the total energy per atom of the Li-metal and the Faraday constant (= 1), respectively. The calculated AIV for Sn4/(8,8) CNT, Li2Sn4/(8,8) CNT, Li4Sn4/(8,8) CNT, Li6Sn4/(8,8) CNT, and Li8Sn4/(9,9) CNT systems are ≈0.49, 0.22, 0.19, 0.07, and −0.11 V, respectively. Thus, our results show that the Li–Sn–CNT composite can potentially be used as the anode of Li-ion batteries. Moreover, the volume change during lithiation is observed to be minimal (only 0.33%).

Based on the energetics obtained from the first principles calculations, it was found that the lithiation of the anode during the first cycle of charging happens in three stages. First, the Li ion prefers to alloy with Sn to form the Zintl cluster; second, it forms a composite via cation–π interaction; and third, it intercalates in the pores of the CNT assembly. Now, the composite formation energy of Li–Sn–CNT system is defined as Ec = Eal + Ead. The first term, Eal is estimated from:

 
Eal = E(Sn4) + (Li) − E(LinSn4) (4)
where μ(Li) is the chemical potential of the Li metal, and E(Sn4) and E(LinSn4) are the total energies of the Sn4 and LinSn4 clusters, respectively. Here, we used the chemical potential of Li to calculate Eal as alloying occurs between the Sn4 cluster and the Li metal. The calculated Ec values are shown in Fig. 6 and it can be inferred that up to n = 4, the alloying process dominates, whereas for n > 4, Ead is found to be larger than Eal. Hence, the Sn–Li–carbon composite is formed in this region. To determine the feasibility of the intercalation process, we also plotted the intercalation energy (shown as dotted lines in Fig. 6) of LinSn4 (n = 0, 2, 4, 6) clusters by considering the respective Ec as the origin. From this, it is clearly observed that the value of Ec is always higher than that of the intercalation energy until n = 6; so, the intercalation in the pores of the CNTs assembly dominates only after this limit. Since the Eal and Ead values are comparatively larger than the energetics of the intercalation process, it is very difficult to gain back all the Li ions from the composites for n ≤ 6. Thus, the capacity of the material is expected to fade as observed experimentally.20 However, a constant capacity will be retained afterwards as Ec is saturated beyond n = 6.


image file: c4ra11187g-f6.tif
Fig. 6 The composition energy of Li–Sn–CNT (total) versus number of Li ions (n) is shown along with the Sn–Li alloying energy (Eal), adsorption energy (Ead) and intercalation energy (dotted line).

4. Conclusions

By employing the first principles density functional calculations, the structural stability and electronic properties of LinSn4 (n = 0–10) clusters and cluster loaded CNTs were studied. Among the clusters, the Li4Sn4 Zintl cluster was found to be the most stable as it has tetrahedral symmetry and largest HOMO–LUMO gap. Our calculations revealed that the insertion of the non-lithiated Sn4 cluster into CNT was shown to be unstable due to the negative adsorption energy which is reflected as charge polarization on the CNT as well as the cluster. On the other hand, the lithiated Sn cluster is preferred for insertion into CNT as significant charge transfer occurs from Sn-5p states to C-2p states through Li ions by cation–π interaction, which was confirmed by electronic density of states and band structure calculations. A similar study was repeated for the deposition of clusters onto graphene and the results were almost reproduced with a smaller effect owing to its planar nature. As a result of the strong interaction between the CNT and the lithiated cluster, the number of Li dimers is reduced, as well as, in certain cases, the geometry of SWCNT being deformed from regular circular to elliptic cylinder shape. We also demonstrated that this deformation is controlled by introducing MWCNT. In addition, we also showed that this composite could be used as the anode of Li-ion batteries as the maximum intercalation potential was observed to be ≈0.5 V.

Acknowledgements

We acknowledge T. Prem Kumar for initiating this work. This work is supported by CSIR, India through TAPSUN (NWP-56) project. We also extend it for CSIR – CECRI, CSIR – NCL and CSIR – CMMACS for sharing their HPC facilities.

Notes and references

  1. M. S. Dresselhaus, G. Dresselhaus and P. C. Eklund, Science of Fullerenes and Carbon Nanotubes, Academic press, 1996 Search PubMed.
  2. (a) P. Ball, Nature, 1991, 354(6348), 18 CrossRef; (b) H. P. Schultz, J. Org. Chem., 1965, 30, 1361 CrossRef CAS.
  3. A. K. Geim and K. S. Novoselov, Nat. Mater., 2007, 6, 183 CrossRef CAS PubMed.
  4. M. Endo and H. W. Kroto, J. Phys. Chem., 1992, 96, 6941 CrossRef CAS.
  5. J. B. Goodenough and K.-S. Park, J. Am. Chem. Soc., 2013, 135, 1167 CrossRef CAS PubMed.
  6. C. K. Chan, R. Ruffo, S. S. Hong, R. A. Huggins and Y. Cui, J. Power Sources, 2009, 189, 34 CrossRef CAS PubMed.
  7. H. Wu, G. Zheng, N. Liu, T. J. Carney, Y. Yang and Y. Cui, Nano Lett., 2012, 12, 904 CrossRef CAS PubMed.
  8. Z. Ma, T. Li, Y. L. Huang, J. Liu, Y. Zhou and D. Xueb, RSC Adv., 2013, 3, 7398 RSC.
  9. C. de las Casas and W. Li, J. Power Sources, 2012, 208, 74 CrossRef CAS PubMed.
  10. K.-C. Lee, J. Y. Jung, S.-W. Lee, H.-S. Moon and J.-W. Park, J. Power Sources, 2004, 129, 270 CrossRef CAS PubMed.
  11. E. Yo, J. Kim, E. Hosono, H. Zhou, T. Kudu and I. Honma, Nano Lett., 2008, 8, 2277 CrossRef PubMed.
  12. B. J. Landi, M. J. Ganter, C. D. Cress, R. A. DiLeo and R. P. Raffaelle, Energy Environ. Sci., 2009, 2, 638 CAS.
  13. C.-M. Park, J.-H. Kim, H. Kim and H.-J. Sohn, Chem. Soc. Rev., 2010, 39, 3115 RSC.
  14. T. Zhang, L. J. Fu, J. Gao, Y. P. Wu, R. Holze and H. Q. Wu, J. Power Sources, 2007, 174, 770 CrossRef CAS PubMed.
  15. X. Wang, Z. Wen, Y. Liu and X. Wu, Electrochim. Acta, 2009, 54, 4662 CrossRef CAS PubMed.
  16. P. Meduri, C. Pendyala, V. Kumar, G. U. Sumanasekara and M. K. Sunkara, Nano Lett., 2009, 9, 612 CrossRef CAS PubMed.
  17. S. Scharfe, F. Kraus, S. Stegmaier, A. Schier and T. F. Fassler, Angew. Chem., Int. Ed., 2011, 50, 3630 CrossRef CAS PubMed.
  18. R. Winter, O. Leichtweib, C. Biermann, M.-L. Saboungi, R. L. McGreevy and W. S. Howells, J. Non-Cryst. Solids, 1996, 66, 205 Search PubMed.
  19. M. Winter and J. O. Besenhard, Electrochim. Acta, 1999, 45, 31 CrossRef CAS.
  20. T. P. Kumar, R. Ramesh, Y. Y. Lin and G. T.-K. Fey, Electrochem. Commun., 2004, 6, 520 CrossRef CAS PubMed.
  21. Y. Yu, L. Gu, C. Wang, A. Dhanabalan, P. A. van Aken and J. Maier, Angew. Chem., Int. Ed., 2009, 48, 6485 CrossRef CAS PubMed.
  22. G. Derrien, J. Hassoun, S. Panero and B. Scrosati, Adv. Mater., 2007, 19, 2336 CrossRef CAS.
  23. J. Qin, C. He, Z. Wang, C. Shi and E.-Z. Liu, ACS Nano, 2014, 8, 1728 CrossRef CAS PubMed.
  24. Y. Sun, Q. Wu and G. Shi, Energy Environ. Sci., 2011, 4, 1113 CAS.
  25. G. Wang, B. Wang, X. Wang, J. Park, S. Dou, H. Ahn and K. Kim, J. Mater. Chem., 2009, 19, 8378 RSC.
  26. M. Noh, Y. Kwon, H. Lee, J. Cho, Y. Kim and M. G. Kim, Chem. Mater., 2005, 17, 1926 CrossRef CAS.
  27. Y. Wang, M. Wu, Z. Jiao and J. Y. Lee, Chem. Mater., 2009, 21, 3210 CrossRef CAS.
  28. Y. Yu, L. Gu, C. Zhu, P. A. van Aken and J. Maier, J. Am. Chem. Soc., 2009, 131, 15984 CrossRef CAS PubMed.
  29. Z. Wen, S. Cui, H. Kim, S. Mao, K. Yu, G. Lu, H. Pu, O. Mao and J. Chen, J. Mater. Chem., 2012, 22, 3300 RSC.
  30. Y. Xu, J. Guo and C. Wang, J. Mater. Chem., 2012, 22, 9562 RSC.
  31. M.-F. Ng, J. Zheng and P. Wu, J. Phys. Chem. C, 2010, 114, 8542 CAS.
  32. W. Jiang, W. Zeng, Z. Ma, Y. Pan, J. Lin and C. Lu, RSC Adv., 2014, 4, 41281 RSC.
  33. G. Kresse and D. Joubert, Phys. Rev. B, 1999, 59, 1758 CrossRef CAS.
  34. P. E. Blochl, Phys. Rev. B, 1994, 50, 17953 CrossRef.
  35. J. P. Perdew, J. A. Chevary, S. H. Vosko, K. A. Jackson, M. R. Pederson, D. J. Singh and C. Fiolhais, Phys. Rev. B, 1992, 46, 6671 CrossRef CAS.
  36. R. Pushpa, S. Narasimhan and U. Waghmare, J. Chem. Phys., 2004, 121, 5211 CrossRef CAS PubMed.
  37. W. J. Zheng, O. C. Thomas, J. M. Nilles, K. H. Bowen, A. C. Reber and S. N. Khanna, J. Chem. Phys., 2011, 134, 224307 CrossRef PubMed.
  38. B. Wang, M. J. Stott and J. A. Alonso, Phys. Rev. B, 2002, 65, 045410 CrossRef.
  39. M. S. Lee, D. G. Kanhere and K. Joshi, Phys. Rev. B, 2005, 72, 015201 Search PubMed.
  40. X. Fan, W. T. Zheng, J.-L. Kuo and D. J. Singh, ACS Appl. Mater. Interfaces, 2013, 5, 7793 CAS.
  41. K. Rana, G. K. Dogu, H. S. Sen, C. Boothroyd, O. Gulseren and E. Bengu, J. Phys. Chem. C, 2012, 116, 11364 CAS.
  42. A. S. Mahadevi and G. N. Sastry, Chem. Rev., 2012, 113, 2100 CrossRef PubMed.
  43. G. Henkelman, A. Arnaldsson and H. Jonsson, Comput. Mater. Sci., 2006, 36, 254 CrossRef PubMed.
  44. K. Persson, Y. Hinuma, Y. S. Meng, A. V. der Ven and G. Ceder, Phys. Rev. B, 2010, 82, 125416 CrossRef.

Footnote

Electronic supplementary information (ESI) available: [1] Kohn–Sham energy levels of various clusters, [2] adsorption energies (Ead) of LinSn4 clusters with various sized armchair CNT, [3] optimized structure of Li6Sn4/(10,10) CNT system, [4] charge transfer diagram of Li8Sn4/(12,6) and Li8Sn4/(15,0) systems, [5] electronic density of states of Li8Sn4/(12,6) and Li8Sn4/(15,0) systems, [6] charge transfer diagram of Li4Sn4/Gr, Li6Sn4/Gr and Li8Sn4/Gr systems, [7] optimized structures of Li8Sn4/(9,9) CNT and Li8Sn4/double walled CNTs (DWCNT) systems and [8] Li ion intercalation in hexagonal assembly of Li6Sn4/(8,8) CNT and Li8Sn4/(9,9) CNT systems. See DOI: 10.1039/c4ra11187g

This journal is © The Royal Society of Chemistry 2015
Click here to see how this site uses Cookies. View our privacy policy here.