Non-magnetic doping induced a high spin-filter efficiency and large spin Seebeck effect in zigzag graphene nanoribbons
Received
10th August 2013
, Accepted 7th October 2013
First published on 8th October 2013
Abstract
Based on nonequilibrium Green's functions (NGF) and density-functional theory (DFT), we investigate the magnetotransport properties and magnetothermoelectric effects in zigzag graphene nanoribbons (ZGNRs) with non-magnetic doping on the double ribbon edges. One of the carbon atoms without hydrogen saturation in each ribbon edge is replaced by one boron (B) or one nitrogen (N) atom. Compared with boron–boron (BB) and nitrogen–nitrogen (NN) double-edge doping, the boron–nitrogen (BN) double-edge doping induces a perfect spin-filter effect with 100% negative spin polarization at the Fermi level. Moreover, we find that the thermoelectric effect can be enhanced by the double-edge doping. Interestingly, the spin Seebeck effect in NN- and BN-doped ZGNRs becomes comparable with the charge Seebeck effect and even larger than it. These results originate from the spin-dependent transmission node near the Fermi level induced by the non-magnetic doping. These findings strongly suggest that the double-edge-doped ZGNRs are promising materials for spintronics and thermo-spintronics.
Introduction
Graphene was first discovered experimentally in 2004.1 It has a typical two-dimensional honeycomb structure made up entirely of carbon atoms (1s22s22p2). Owing to its superior physical properties and potential applications in electronics and spintronics, graphene has attracted considerable attention.2,3 Recently, some experimental studies have shown that graphene nanoribbons (GNRs) can be produced by patterning the graphene sheet into ribbons or by chemical methods.4,5 Jiao et al. proposed an approach to producing GNRs with smooth edges and a narrow width by unzipping multi-walled carbon nanotubes.6 According to the edge characteristics, the GNRs are divided into armchair graphene nanoribbons (AGNRs) and zigzag graphene nanoribbons (ZGNRs). Many theoretical and experimental studies have shown that AGNRs are semiconducting with the energy gap scaling as the inverse of their width,7,8 and ZGNRs are magnetic semiconductors with spin-polarized edge states.9,10 The energy gap of ZGNRs is attributed to an unusual antiferromagnetic coupling between opposite-edge carbon atoms and its size is also inversely proportional to its width.
In 2006 Son et al. provided a pioneering idea of inducing a half-metallic behavior in zigzag GNRs by an external electric field.9 In addition, chemical dopants can also change the electronic and magnetic properties of ZGNRs. For instance, Dutta et al. performed a first-principles calculation to explore the electronic properties of ZGNRs modified with chemical dopants, boron (1s22s22p1) and nitrogen (1s22s22p3). When two middle zigzag carbon chains are replaced by boron–nitrogen (BN) chains, the modified ZGNR exhibits both semiconducting and intrinsic half-metallic behaviors under the influence of an external electric field.11 Lin et al. showed that a topological line defect could also yield a half-metallic behavior when it is close to one edge.12 A single-spin negative differential resistance in beryllium (Be)-edge-doped ZGNRs has also been reported,13 indicating that one single-spin current has a negative differential resistance whereas the opposite single-spin current shows a monotonic behavior as the external bias increases. In ref. 14, the authors investigated the spin-polarized transport properties of hybrid graphene nanoribbons, and showed a remarkable width-dependent magnetic behavior. Zheng et al. investigated the effects of nitrogen (N) and boron (B) impurities on the electron transmission function of ZGNRs by first principles calculations.15 Their results showed that bound states and quasibound states can be introduced by a B/N impurity, resulting in many interesting transport properties. Biel et al. presented a first-principles calculation of the quantum transport in chemically doped graphene nanoribbons by B and N impurities.16 A resonant backscattering was yielded, and they further found an unusual acceptor–donor transition in ZGNRs.
Thermoelectric effects convert heat energy into electric energy or vice versa. Some experiments have demonstrated the capability of measuring the Seebeck coefficient at atomic and molecular junctions.17,18 More recently, using the imaging capability of the scanning tunneling microscope (STM), Evangeli et al. provided an approach to measuring continuously and simultaneously the thermopower and conductance of C60 molecular junctions.19 The surface reconstruction can occur at the interface of C60 absorbed on the Cu(111) substrate at room temperature. Then Hsu et al. investigated the effects of such reconstruction on the thermopower of C60 molecular junctions.20 Widawsky et al. investigated the length-dependent conductance and thermopower of Au–C bonded single-molecule junctions. A nonlinear relationship between the thermopower and the molecular length was shown, while the conductance decreases exponentially as the molecular length increases.21 Encouragingly, Uchida et al. measured a spin voltage from a temperature gradient in a metallic magnet (Ni81Fe19) with the help of a latest developed spin-detection technique.22 This phenomenon is called the spin Seebeck effect and provides a method to produce a pure spin current by using the temperature field. Until now this pioneering experiment has inspired many theoretical and experimental studies on the spin thermoelectric effect in various systems.23–27 The new branch of spintronics can be named as “thermo-spintronics” or “spin-caloritronics”, and focuses on studies of the manipulation of the spin degree of freedom by temperature fields. Recently, the spin-dependent thermoelectric effect in spin valves, consisting of ZGNRs with different configurations, has been investigated by first-principles calculations combined with non-equilibrium Green's functions.28 Moreover, a mean-field Hubbard model has also been employed to investigate thermally driven spin-polarized transport properties of a locally gated ZGNR. The thermal magnetoresistance can reach 105% at small gate voltages.29
In this paper, we investigate simultaneously the spin-filter effect and thermoelectric performance of ZGNRs with chemical dopants. The proposed molecular devices are shown in Fig. 1, which are divided into three regions: left electrode, scattering region, and right electrode. One of the carbon atoms without hydrogen saturation in each ribbon edge is replaced by one B or N atom. It is found that the BN double-edge-doped ZGNR shows a half-metallic behavior with 100% negative spin polarization at the Fermi level. Moreover, we also find that the spin Seebeck effect can become comparable with the charge Seebeck effect in these non-magnetic doped ZGNRs and even larger than it. The findings here can be interpreted as the results of the spin-dependent transmission node near the Fermi level.
 |
| Fig. 1 Schematic of two-probe systems for four non-magnetic-doped ZGNRs. (a), (b), (c) and (d) are without doping (CC–ZGNR), with boron and boron doping (BB–ZGNR), nitrogen and nitrogen doping (NN–ZGNR), and boron and nitrogen doping (BN–ZGNR) at the two edges of the ZGNRs, respectively. The gray balls are carbon atoms, and the bright balls are hydrogen atoms. The blue and brown balls represent nitrogen and boron atoms, respectively. | |
Theoretical methods
All ab initio calculations including the structure relaxation, transmission function, band structure, and density of states (DOS) in this work are implemented by using the ATK (Atomistix ToolKit) package, which is based on the density-functional theory (DFT) combined with the nonequilibrium Green's function (NGF) method.30,31 The electron wave function is expanded in a basis set of double zeta orbitals plus one polarization orbital (DZP). To separate ZGNRs in neighbor supercells and ensure the suppression of the coupling between them, a vacuum region of width 15 Å is adopted. All structures in Fig. 1 are optimized until the inter-atomic forces are less than 0.05 eV Å−1. The energy cutoff is chosen to be 150 Ry and the size of the mesh grid in k space for the electrode parts is 1 × 1 × 100. In the ATK package the spin-dependent transmission function τσ(ε) of electrons with a spin index σ and energy ε is | τσ(ε) = Tr[ΓLσ(ε)Grσ(ε)ΓRσ(ε)Gaσ(ε)], | (1) |
where ΓL(R)σ(ε) is the line-width function of electrons with a spin index σ and energy ε describing the coupling between the scattering region and the left (right) electrode. The retarded Green's function is obtained by Grσ(ε) = [εI − H + i(ΓLσ + ΓRσ)/2]−1, where I is the unit matrix and H is the Hamiltonian of the scattering region. The advanced Green's function Gaσ(ε) is obtained from the relationship Gaσ = [Grσ]†. The spin polarization at the Fermi level EF is determined by |  | (2) |
where EF is the Fermi level of the electrodes in the absence of the external voltage. In the linear response region the spin-dependent Seebeck coefficient can be obtained by32,33 |  | (3) |
where Kνσ(EF,T) = −∫dε{∂f(ε,EF,T)/∂ε}(ε − EF)ντσ(ε)(ν = 0,1). T and e are the electrode's temperature and electron charge, respectively. Furthermore, f(ε,EF,T) = [1 + e(ε−EF)/kBT]−1 is the Fermi–Dirac distribution function in the electrodes and kB is the Boltzmann constant. If τσ(ε) has a smooth change near the Fermi level, Sσ in the low-temperature region reduces to34 |  | (4) |
The above equation clearly shows that Sσ is not only related to the magnitude of τσ(ε) but also related to the slope of τσ(ε) at the Fermi level. Therefore, the study of Sσ can provide more information than current–voltage investigations. But if τσ(ε) has a huge change near the Fermi level, eqn (4) is no longer applicable even in the low-temperature region. For instance, a symmetrical transmission node at the Fermi energy yields an obvious enhancement of single-molecule thermoelectric effects on the two sides of the node.35
Results and discussions
In this work, we only consider the ferromagnetic state, which plays a very important role in the spin-polarized transport properties near the Fermi level of the ZGNRs.36,37 Moreover, the ZGNRs can be derived from an antiferromagnetic ground state to a ferromagnetic state by a magnetic field. Now we start with the study of the spin-filter effect for the four two-probe systems based on ZGNRs shown in Fig. 1. In the left column of Fig. 2, we show the transmission function τσ as a function of the electron energy. For a perfect ZGNR without any doping, each spin channel at the Fermi level is completely open. In this case the transmission function τσ is 1 and the spin polarization ζ is zero (see Fig. 2(a) and Table 1). It is interesting to note that the spin-up and spin-down transmission functions exhibit different sensitivities to the double-edge doping. In the BB–ZGNR the spin-up transmission function at the Fermi level is suppressed, while the spin-down one almost remains unchanged (see Fig. 2(b) and Table 1). In contrast, the spin-up transmission function at the Fermi level in the NN–ZGNR does not change, while the spin-down one is suppressed as shown in Fig. 2(c). Therefore, the spin polarization can be modulated from negative to positive when two B atoms are replaced by two N atoms. If one of the two B atoms is replaced by one N atom, we find that a spin-up transmission node is exactly located at the Fermi level. The spin-down transmission function τ↓ is also slightly suppressed (see Fig. 2 (d)). The corresponding spin polarizations for the BB–, NN–, and BN–ZGNRs are shown in Table 1. In particular, it should be noted that the value of ζ for the BN–ZGNR approaches −100%, indicating the coexistence of a metallic state for spin-down electrons and an insulating state for spin-up electrons at the Fermi level. These transmission nodes near the Fermi level are attributed to the appearance of some quantum states close to the Fermi level induced by the double-edge doping. In the DOS spectrum (see the middle column in Fig. 2) and band structure (see the right column in Fig. 2), we see clearly that these quantum states indeed appear near the Fermi level. Our results are similar to those reported in ref. 38. The locations of these quantum states are marked by blue arrows, as shown in Fig. 2. For the BB–ZGNR, one spin-up quantum state near the Fermi level appears, and a spin-down quantum state for the NN–ZGNR appears near the Fermi level. But for the BN–ZGNR, the spin-up and spin-down quantum states appear concurrently at different locations near the Fermi level, resulting in two transmission nodes with different spin components.
 |
| Fig. 2 Transmission functions (left column), density of states (middle column) and band structure (right column) for the four ZGNRs. The band structure is obtained by taking the center region as the periodic structure. From top to bottom, the structure is CC–ZGNR, BB–ZGNR, NN–ZGNR, and BN–ZGNR, respectively. The black solid lines denote spin-up components and the read dashed lines for the spin-down ones. The blue arrows show the locations of the single-spin quantum states induced by the double-edge doping. | |
Table 1 Transmission function and spin polarization at the Fermi level EF
Molecule |
τ
↑
|
τ
↓
|
ζ
|
CC–ZGRN |
1 |
1 |
0% |
BB–ZGRN |
0.1638 |
0.9993 |
−71.8% |
NN–ZGRN |
0.9839 |
0.5318 |
29.8% |
BN–ZGRN |
0.0009 |
0.7122 |
−99.7% |
To further understand the spin anisotropy at the Fermi level, we show the corresponding spin-dependent local density of state (LDOS) in Fig. 3. For the pure ribbon (CC–ZGNR), the spin-up and spin-down electronic states are uniformly distributed along the z direction in the scattering region. Moreover, quantum size effects make the electrons pile up on the two edges of the CC–ZGNR. When one of the carbon atoms without hydrogen saturation in each ribbon edge is replaced by one B atom (see Fig. 1(b)), the spin-up electronic states along the z direction become nonuniform in the scattering region and the corresponding LDOS close to left and right electrodes is obviously suppressed. However, the spin-down LDOS at the Fermi level is almost unchanged. Consequently, τ↑ at the Fermi level is suppressed, while τ↓ is retained. In the NN–ZGNR the situation is inverse. For example, the spin-down LDOS is obviously suppressed in the NN scattering region and the region close to electrodes, while the spin-up LDOS is retained. This fact results in 29.8% spin polarization. A more intriguing phenomenon in the BN–ZGNR is that the spin-up electronic state at the Fermi level is completely localized in the scattering region and the corresponding LDOS is mainly concentrated in the center region of one edge. The spin-down LDOS at the Fermi level is slightly suppressed in the BN scattering region and the region close to electrodes. These facts induce a transmission node for spin-up electrons at the Fermi level (τ↑(EF) ≃ 0), and there is a finite value for the spin-down transmission function (τ↓(EF) = 0.7). Accordingly, the spin polarization at the Fermi level approaches −100%. An ideal spin-filter effect is achieved in the BN–ZGNR.
 |
| Fig. 3 Spin-dependent LDOS of four two-probe systems at the Fermi level. | |
The study of the Seebeck effect can provide an effective avenue to explore the electronic structures of the devices. The sign of the Seebeck coefficient is related to its type. For example, a negative Seebeck coefficient suggests an n-type device, while a positive Seebeck coefficient corresponds to a p-type device. In the left panel of Fig. 4 we plot Sσ as a function of the temperature for the CC–, BB–, NN–, and BN–ZGNRs, respectively. The spin Seebeck coefficient SS(=(S↑ − S↓)/2) and charge Seebeck coefficient SC(=(S↑ + S↓)/2) versus the temperature are plotted in the right panel of Fig. 4. In the CC–ZGNR, τσ, without any change near the Fermi level, induces a relatively small Seebeck effect in the low-temperature region. But when T is larger than 200 K, Sσ is slightly enhanced. This result originates from the contributions of side electronic states as the temperature increases, and one should note a two-peak structure near the Fermi level in the transmission spectrum of the CC–ZGNR (see Fig. 2(a)). For the BB–ZGNR and NN–ZGNR, we find that the sign of Sσ is negative, which indicates that these devices are n-type. The results can be explained by eqn (4). The slope of τσ of the BB–ZGNR and NN–ZGNR at the Fermi level is positive, which results in a negative value of Sσ due to Sσ ∝ − τ′σ(ε)/τσ(ε)|ε=EF. For the BB–ZGNR, we have τ′↑(EF) > τ′↓(EF) and τ↑(EF) < τ↓(EF), while the relationships for the NN–ZGNR are τ′↑(EF) < τ′↓(EF) and τ↑(EF) > τ↓(EF). Therefore, |S↑| is larger than |S↓| in the BB–ZGNR, while |S↓| is larger than |S↑| for the NN–ZGNR. We also note that τ↓ at the Fermi level for the NN–ZGNR has a larger value than τ↑ at the Fermi level for the BB–ZGNR. This fact results in a larger value of S↑ for the BB–ZGNR than S↓ for the NN–ZGNR. More interestingly, S↑ for the BN–ZGNR shows an obvious nonmonotonic behavior as the temperature increases. In the low-temperature region S↑ has a positive maximum at about 25 K. The sign of S↑ changes from negative to positive as the temperature increases, which indicates that the type of devices can be manipulated by the temperature. This result is ascribed to an asymmetrical transmission node for spin-up electrons at the Fermi level. For T ≤ 100 K the transmission function τ↑ below the Fermi level plays a more important role than τ↑ above the Fermi level. This fact results in a positive value of S↑. As the temperature increases, τ↓ above the Fermi level will contribute more and more to S↑. Therefore, S↑ becomes negative for T ≥ 100 K. In addition, it is interesting that SC and SS become comparable in the high-temperature region for the NN–ZGNR and in the low-temeprature region for the BN–ZGNR. SS is even larger than SC for the BN–ZGNR (see the inset of Fig. 4(h)).
 |
| Fig. 4 Spin-dependent Seebeck coefficient Sσ as a function of temperature for (a) CC–ZGNR, (c) BB–ZGNR, (e) NN–ZGNR, and (g) BN–ZGNR. The corresponding spin Seebeck coefficient SS and charge Seebeck coefficient SC as functions of temperature for (b) CC–ZGNR, (d) BB–ZGNR, (f) NN–ZGNR, and (h) BN–ZGNR. The inset in (h) magnifies the region 0 ≤ T ≤ 50 K. | |
Conclusions
In summary, we have studied the effects of non-magnetic doping in the ribbon edges on magneto-transport properties and magneto-thermoelectric effects of ZGNRs. ZGNR-based devices with a high spin-filter efficiency and large spin thermoelectric effect have been proposed. The spin-dependent transmission function near the Fermi level can be effectively modulated by double-edge doping; as a result, the spin-filter efficiency and spin thermoelectric effect are enhanced. When one carbon atom without hydrogen saturation in each edge is replaced simultaneously by one boron and one nitrogen atom, an ideal half-metallic behavior is found. The reason is that a spin-up localized state emerges at the Fermi level and the spin-up channel is completely closed. More interestingly, the spin Seebeck effect SS is comparable to and even larger than the charge Seebeck effect SC for the BN–ZGNR. In general, these findings strongly suggest that simple chemical dopants can induce a perfect spin-filter effect and large spin Seebeck effect in ZGNRs.
Acknowledgements
The authors thank the support of the National Natural Science Foundation of China (NSFC) under Grants No. 11247028, 61306122, 61274101 and 91121021. The work was also sponsored by Qing Lan Project.
References
- K. S. Novoselov, A. K. Geim, S. V. Morozov, D. Jiang, Y. Zhang, S. V. Dubonos, I. V. Grigorieva and A. A. Firsov, Science, 2004, 306, 666 CrossRef CAS PubMed.
- A. K. Geim, Rev. Mod. Phys., 2011, 83, 851 CrossRef CAS.
- K. S. Novoselov, Rev. Mod. Phys., 2011, 83, 837 CrossRef CAS.
- L. Tapaseto, G. Dobrik, P. Lambin and L. P. Biró, Nat. Nanotechnol., 2008, 3, 397 CrossRef PubMed.
- A. K. Geim and K. S. Novoselov, Nat. Mater., 2007, 6, 183 CrossRef CAS PubMed.
- L. Y. Jiao, L. Zhang, X. R. Wang, G. Diankov and H. J. Dai, Nature, 2009, 458, 877 CrossRef CAS PubMed.
- V. Barone, O. Hod and G. E. Scuseria, Nano Lett., 2006, 6, 2748 CrossRef CAS PubMed.
- M. Y. Han, B. Özyilmaz, Y. B. Zhang and P. Kim, Phys. Rev. Lett., 2007, 98, 206805 CrossRef.
- Y. W. Son, M. L. Cohen and S. G. Louie, Phys. Rev. Lett., 2006, 97, 216803 CrossRef.
- J. Jung, T. Pereg-Barnea and A. H. MacDonald, Phys. Rev. Lett., 2009, 102, 227205 CrossRef CAS.
- S. Dutta, A. K. Manna and S. K. Pati, Phys. Rev. Lett., 2009, 102, 096601 CrossRef.
- X. Q. Lin and J. Ni, Phys. Rev. B: Condens. Matter Mater. Phys., 2011, 84, 075461 CrossRef.
- T. T. Wu, X. F. Wang, M. X. Zhai, H. Liu, L. P. Zhou and Y. J. Jiang, Appl. Phys. Lett., 2012, 100, 052112 CrossRef.
- A. R. Botello-Méndez, E. Cruz-Silva, F. López-Urias, B. G. Sumpter, V. Meunier, M. Terrones and H. Terrones, ACS Nano, 2009, 3, 3606 CrossRef PubMed.
- X. H. Zheng, I. Rungger, Z. Zeng and S. Sanvito, Phys. Rev. B: Condens. Matter Mater. Phys., 2009, 80, 235426 CrossRef.
- B. Biel, X. Blase, F. Triozon and S. Roche, Phys. Rev. Lett., 2009, 102, 096803 CrossRef.
- P. Reddy, S. Y. Jang, R. A. Segalman and A. Majumdar, Science, 2007, 315, 1568 CrossRef CAS PubMed.
- K. Baheti, J. A. Malen, P. Doak, P. Reddy, S. Y. Jang, T. D. Tilley, A. Majumdar and R. A. Segalman, Nano Lett., 2008, 8, 715 CrossRef CAS PubMed.
- C. Evangeli, K. Gillemot, E. leary, M. T. González, G. Rubio-Bollinger, C. J. Lambert and N. Agraït, Nano Lett., 2013, 13, 2141 CrossRef CAS PubMed.
- B. C. Hsu, C. Y. Lin, Y. S. Hseieh and Y. C. Chen, Appl. Phys. Lett., 2012, 101, 243103 CrossRef.
- J. R. Widawsky, W. Chen, H. Vázquez, T. Kim, R. Breslow, M. S. Hybertsen and L. Venkataraman, Nano Lett., 2013, 13, 2141 CrossRef PubMed.
- K. Uchida, S. Takahashi, K. Harii, J. Ieda, W. Koshibae, K. Ando, S. Maekawa and E. Saitoh, Nature, 2008, 455, 778 CrossRef CAS PubMed.
- S. Bosu, Y. Sakuraba, K. Uchida, K. Saito, T. Ota, E. Saitoh and K. Takanashi, Phys. Rev. B: Condens. Matter Mater. Phys., 2008, 83, 224401 CrossRef.
- C. M. Jaworski, J. Yang, S. Mack, D. D. Awschalom, J. P. Heremans and R. C. Myers, Nat. Mater., 2010, 9, 898 CrossRef CAS PubMed.
- K. Uchida, H. Adachi, T. An, T. Ota, M. Toda, B. Hillebrands, S. Maekawa and E. Saitoh, Nat. Mater., 2011, 10, 737 CrossRef CAS PubMed.
- H. Adachi, J. Ohe, S. Takahashi and S. Maekawa, Phys. Rev. B: Condens. Matter Mater. Phys., 2011, 83, 094410 CrossRef.
- Y. Dubi and M. Di Ventra, Rev. Mod. Phys., 2011, 83, 131 CrossRef CAS.
- M. G. Zeng, W. Huang and G. C. Liang, Nanoscale, 2013, 5, 200 RSC.
- Z. Y. Zhao, X. C. Zhai and G. J. Jin, Appl. Phys. Lett., 2012, 101, 083117 CrossRef.
- J. Taylor, H. Guo and J. Wang, Phys. Rev. B: Condens. Matter Mater. Phys., 2008, 63, 245407 CrossRef.
- Atomistix ToolKit: Manual, Version 2008. 10.
- Y. Dubi and M. Di Ventra, Phys. Rev. B: Condens. Matter Mater. Phys., 2009, 79, 081302(R) Search PubMed.
- Y. S. Liu, X. F. Yang, F. Chi, M. S. Si and Y. Guo, Appl. Phys. Lett., 2012, 101, 213109 CrossRef.
- Y. S. Liu, Y. R. Chen and Y. C. Chen, ACS Nano, 2009, 3, 3497 CrossRef CAS PubMed.
- J. P. Bergfield and C. A. Stafford, Nano Lett., 2009, 9, 3072 CrossRef CAS PubMed.
- T. B. Martins, R. H. Miwa, A. J. R. da Silva and A. Fazzio, Phys. Rev. Lett., 2007, 98, 196803 CrossRef CAS.
- C. Q. Qu, C. Y. Wang, L. Qiao, S. S. Yu and H. B. Li, Chem. Phys. Lett., 2013, 578, 97 CrossRef CAS PubMed.
- E. Cruz-Silva, Z. M. Barneett, B. G. Sumpter and V. Meunier, Phys. Rev. B: Condens. Matter Mater. Phys., 2011, 83, 155445 CrossRef.
|
This journal is © The Royal Society of Chemistry 2013 |
Click here to see how this site uses Cookies. View our privacy policy here.