This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Edge and sublayer degrees of freedom for phosphorene nanoribbons with twofold-degenerate edge bands via electric field

Yi Ren    Xiaoying Zhou    Guanghui Zhou [email protected] Department of Physics, Key Laboratory for Low-Dimensional Quantum Structures and Quantum Control (Ministry of Education), and Synergetic Innovation Center for Quantum Effects and Applications, Hunan Normal University, Changsha 410081, China
Abstract

For the pristine phosphorene nanoribbons (PNRs) with edge states, there exist two categories of edge bands near the Fermi energy (EFE_{F}), i.e., the shuttle-shaped twofold-degenerate and the near-flat simple degenerate edge bands. However, the usual experimental measurement may not distinguish the difference between the two categories of edge bands. Here we study the varying rule for the edge bands of PNRs under an external electrostatic field. By using the KWANT code based on the tight-binding approach, we find that the twofold-degenerate edge bands can be divided into two separated shuttles until the degeneracy is completely removed and a gap near EFE_{F} is opened under a sufficiently strong in-plane electric field. Importantly, each shuttle from the ribbon upper or lower edge outmost atoms is identified according to the local density of states. However, under a small off-plane field the shuttle-shaped bands are easily induced into two near-flat bands contributed from the edge atoms of the top and bottom sublayers, respectively. The evidence provides the edge and sublayer degrees of freedom (DOF) for the PNRs with shuttle-shaped edge bands, of which is obviously different from another category PNRs intrinsically with near-flat edge bands. This is because that the former category of ribbons solely have four zigzag-like atomic configurations at the edges in each unit cell, which also results in that the property is robust against the point defect in the ribbon center area. As an application, furthermore, based on this issue we propose a homogenous junction of a shuttle-edge-band PNR attached by two electric gates. Interestingly, the transport property of the junction with field manipulation well reflects the characteristics of the two DOFs. These findings may provide a further understanding on PNRs and initiate new developments in PNR-based electronics.

pacs:
73.22.-f, 73.63.-b, 68.65.-k

I Introduction

The internal degree of freedom (DOF) of electrons in nanostructures is an important issue of modern condensed matter physics. In addition to the charge and spin DOFs, other ones have also been discussed. For example, in multilayer graphene Novoselov and transition-metal dichalcogenides (TMDs) Mak , both the layer and valley DOFs are presented Jose ; Xu . In addition, new DOF may appear upon tailoring a two-dimensional (2D) material into a nanoribbon. The edge and layer, for instance, have been regarded as tunable equivalents of the spin-one-half DOF in bilayer phosphorene nanoribbons (PNRs) with zigzag-edge Soleimanikahnoj . However, here we treat the edge and sublayer as two DOFs in a recently revealed monolayer slope-edged PNR (sPNR) with twofold-degenerate edge bands Ren .

Phosphorene is a few- or mono-layer black phosphorus (BP) where P atoms arranged in the top and bottom sublayers of a puckered honeycomb lattice Li ; Ling . Inside phosphorene, each P atom is covalently bonded with three adjacent atoms to form a puckered honeycomb structure due to the sp3 hybridization HanCQ ; Carvalho . This promising new 2D material, in the sense of applications in nano-electronics, can be exfoliated from bulk BP due to the weak interlayer Van der Waals interaction and possesses a direct band gap of 0.3 eV Li ; HanCQ . This direct gap increases up to \sim2 eV as its thickness decreases from bulk to monolayer Li1 . In the aspect of application, the field-effect transistor (FET) Li ; Steven ; Buscema ; XiaF based on phosphorene is found to have an on/off ratio up to 10310^{3} and a high hole carrier mobility to 800 cm2/Vs Sherman . Further, arising from the low symmetric and highly anisotropic structure, phosphorene owns strongly anisotropic electrical, thermal and optical properties, which may open up possibilities for conceptually new devices LiuH ; HanCQ ; LiPK ; Zhou ; Zhou1 ; LiuP .

On the other hand, nanoribbons can offer better tunability in electronic structures because of the quantum confinement and edge morphological influence. Tailoring a 2D phosphorene sheet along the conventional zigzag and armchair directions have been experimentally realized Paul ; Mitchell . Hence the zigzag PNRs (zPNRs) with significant edge states and armchair PNRs (aPNRs) with a direct band-gap have been extensively studied Carvalho ; Tran ; Ezawa ; ZhangR ; Ajanta , and their skewed or beard counterparts have also been further reported Ezawa ; Marko . In general, the edge states projecting to the outermost atoms of a ribbon in real space are near the Fermi level EFE_{F} Ezawa ; ZhouBL . They have been extensively studied for graphene and MoS2 nanoribbons Nakada ; Bollinger . However, a zPNR has two near-degenerate edge bands closing to EFE_{F}, which are respectively contributed by the atoms of the two edges. And the properties of these important edge states have been recognized ZhouBL ; Sousa ; SK ; Ren1 ; Sisakht ; Soleimanikahnoj . Moreover, we may cut a phosphorene sheet so that the zigzag (armchair) direction intersects the puckered ridges under an chiral angle other than 0º (90º), resulting in PNRs with other possible edge geometries Ren . These ribbons can be classified into two types, one type with edge states including zPNR and the other without edge states including aPNR. In specification, from our previous definition of the chiral vector 𝑻\bm{T}=m𝒂1/2m\bm{a}_{1}/2+n𝒂2/2n\bm{a}_{2}/2 for the crystallographic characterizations on planar phosphorene with chiral numbers (m,nm,n), the cases of m+n=even integer have defined the all possible planar crystal directions Ren . Hence a ribbon can be generally denoted as (m,nm,n)PNR. Furthermore, according to the edge atomic arrangement (morphology), PNRs with edge states can be further divided into two categories. When both m and n are odd, the outermost edge atoms of a PNR are alternately located at the two sublayers, resulting in the shuttle-shaped twofold-degenerate edge bands near EFE_{F}, such as (1,3)PNR and (3,1)PNR Ren ; Marko ; Ashwin ; Liu . While for both even m and n, the outermost atoms are located at the same sublayer and the ribbon only has a near-flat degenerate edge band, such as (2,4)PNR and (4,2)PNR Ren . And some of the typical ribbons with these two categories of edge states have already been observed in experiments Lee ; Liang . However, the electronic property may be significantly different from each other between the two categories. And there are few studies on the PNRs with shuttle-shaped twofold-degenerate edge bands Marko . Therefore, it is essentially demanded to explore the microscopic origin of the twofold-degenerate edge bands, especially the transport property. Meanwhile, the defects in PNR samples, such as monatomic vacancies, are inevitable in experiments Kiraly . It is also important to understand the defect effect on the electronic and transport properties Lill .

In this paper we select two sPNRs, (1,3)PNR and (2,4)PNR, as the exemplary ribbons belonging to the categories with twofold-degenerate and near-flat degenerate edge bands, respectively. By using the KWANT software within the framework of tight-binding method, we find that the two shuttle-shaped twofold-degenerate edge bands of (1,3)PNR are separated until the degeneracy is removed and a gap near EFE_{F} is opened under a sufficiently strong in-plane electric field. And each shuttle contributed from the outmost atoms of the ribbon upper or lower edge are identified according to the local density of states (LDOS). However, under a small off-plane field the shuttle-shape bands are easily separated into two degenerated near-flat bands contributed from the edge atoms of the top or bottom sublayer. The edge band variation with external field for this category is completely different from that of (2,4)PNR belonging to the previously reported zPNR category. This is because a (1,3)PNR has four zigzag atomic configurations on the upper and lower edges and the degenerate bands are from the outermost atoms in the same sublayer or different upper-lower edge. This allows the two DOFs to be distinguished and regulated by applying electric field along different directions. Further, based on this issue we propose a (1,3)PNR homogenous junction attached by electric gates. Interestingly, the transport property of the junction with field manipulation well reflects the characteristics of the two DOFs for (1,3)PNR category. In addition, the defect effect on the transport property is also discussed. The conclusion is that both DOFs are robust against the defect in the center area of the ribbon, but the sublayer DOF is more effective to resist the edge vacancy than the edge DOF. These results may provide a further understanding on PNRs and initiate new developments in PNR-based electronics.

Refer to caption
Figure 1: The schematic illustrations of (a) (1,3)PNR and (b) (2,4)PNR, where the red dashed line parallelogram in each ribbon indicates its minimum periodical supercell, the chiral vector T is illustrated by the red solid arrow along the edge, 𝒂1\bm{a}_{1} and 𝒂2\bm{a}_{2} denote the primitive vectors, and t1t_{1}-t5t_{5} are the five effective hoping parameters. The distance between atomic rows for the two sPNRs is 0.9 Å and 0.53 Å, respectively. And the ribbon end side views in (a) and (b) indicate the top/bottom sublayer by the red/blue color. (c) The sketch of the pp-nn junction composed of a (1,3)PNR where the different color shade in two sides implies the top and bottom back-gates which adjust the Fermi level, and the green thick arrows indicate the in-plane transverse and off-plane vertical electric fields, respectively.

The organization of the paper is as follows. First, we classify the categories of the PNRs with edge states, and then analyze their edge atomic arrangements and unit cell choice In Sec. II, we present the model description and the details on the calculations. Then in Sec. III, we demonstrate the the edge state bands of the two exemplary sPNRs blong to different categories and the edge band variations under in- and off-plane electric fields, respectively. Hence according to the coexistence of two (upper or lower edge and sublayer) DOFs in monolayer (1,3)PNR, with which a pp-nn junction is proposed and the the important transport property is discussed. Finally, the conclusion is briefly drawn in Sec. IV.

II Model Description and Method

The schematic illustrations of the considered (1,3)PNR and (2,4)PNR are shown in Figs. 1(a) and 1(b), respectively. The red dashed parallelograms are the minimum supercells, where the red solid arrows are the chiral vectors indicating the crystal direction of the ribbon edges Ren . The integer NN denotes the number of atomic chains (blue dashed lines) across the ribbon width. Further, for (1,3)PNR shown in Fig. 1(a), the red and blue outmost edge atoms are arranged alternately on the two sublayers (see the side view), as a consequence the supercell width must be the twice of \mid𝑻\bm{T}\mid. In contrast, the outmost edge atoms for (2,4)PNR are always located at one sublayer and the supercell width is equal to \mid𝑻\bm{T}\mid. Moreover, as shown in Fig. 1(c), we propose a homogenous pp-nn junction based on a (1,3)PNR, where the blue and orange background imply the positive and negative potentials provided by back-gate electric gates, respectively. And the thick green arrows in upper and lower panels indicate the in-plane transverse electric field along the yy-direction and the off-plane vertical one along the zz-direction, respectively. The stability of PNRs have been analyzed in our previous work Ren . Since the \mid𝑻\bm{T}\mid of the two selected PNRs here are much less than 20Å, hence they are considered relatively stable and do not appear obvious edge reconstruction.

We use the KWANT code based on tight-binding (TB) Hamiltonian Groth to calculate the electronic energy band for the selected PNRs, and the atomistic quantum-transport simulations for the proposed junction are based on the scattering-matrix method from matching wavefunctions Yongjin ; Zwierzycki ; Tatiane . In comparison with the first-principles calculation Ren1 , this approach can treat large nanostructures matching the usual experimentally reachable sample size up to sub-100 nm scales with better precision without large computational consumption.

The TB model Hamiltonian for phosphorene Rudenko in the presence of an electric field can be described as

H=<i,j>tijcicj+(elεy/z+Ui)icici,H=\sum_{<i,j>}t_{ij}c_{i}^{\dagger}c_{j}+(el\varepsilon_{y/z}+U_{i})\sum_{i}c_{i}^{\dagger}c_{i}\;, (1)

where ci(cj)c_{i}^{\dagger}(c_{j}) is the creation (annihilation) operator of electron at site i(j)i(j), the summation i,j\langle i,j\rangle means over all the neighboring atomic sites with hopping integrals tijt_{ij}, ll is the component of the atomic position from the selected origin along the electric field direction, εy/z\varepsilon_{y/z} is the strength of the field along the y/zy/z direction as shown in Fig. 1(c), and Ui is the impurity-induced potential if available. It has been shown that five hopping parameters [see Fig. 1(b)] are enough to describe the electronic band structure of phosphorene. The values of these hopping integrals suggested in the previous studies Rudenko are t1t_{1}=-1.220 eV, t2t_{2}=3.665 eV, t3t_{3}=-0.205 eV, t4t_{4}=-0.105 eV and t5t_{5}=-0.055 eV. Therefore, by solving the discrete Schrödinger equation corresponding to Hamiltonian (1) on the proper basis for the supercell drawn by the red dashed parallelograms in Fig. 1 and applying the Bloch theorem, the kk-dependent Hamiltonian for a PNR can be written as H(k)H(k)=H0,0+H0,1eika+H0,1eikaH_{0,0}+H_{0,1}e^{ika}+H_{0,1}^{{\dagger}}e^{-ika} in the form of (N×NN^{\prime}\times N^{\prime}) dimensional matrix. Here NN^{\prime} is the number of atoms in the supercell, H0,0H_{0,0} is the matrix of the central cell, H0,1H_{0,1} the coupling matrix with the right-hand adjacent cell, and aa is the length between two nearest-neighbor cells. Diagonalizing this kk-dependent Hamiltonian, we can obtain the band spectrum and the corresponding eigen-wavefunctions. Then we can calculate the LDOS using the following formula

LDOS(E,r)=1c2πn|Ψn(r)|2e(EnE)22c2,\text{LDOS}(E,r)=\frac{1}{c\sqrt{2\pi}}\sum\limits_{n}|\Psi_{n}(r)|^{2}e^{\frac{-(E_{n}-E)^{2}}{2c^{2}}}, (2)

where cc is broadening parameter, Ψn(r)\Psi_{n}(r) and EnE_{n} are the eigen-wavefunction and eigenvalue, respectively, in which nn denotes the energy band index and rr the atom position.

In addition, we mention that the description of the TB model has been effectively used for PNRs with different edge geometries Soleimanikahnoj ; Ren ; LiuP ; ZhangR ; Marko ; ZhouBL . We have verified the validity of TB approach by the first-principles calculation using the Atomistix Toolkit code Smidstrup . The result shows that the edge bands can also be near the Fermi energy and the difference between them is quantitative. As for the influence of spin polarization, it is found that the total energy of (1,3)PNR for the spin polarization case is 8.82 meV lower than that of non-polarized case. And for the (2,4)PNR, the energy difference is only 0.04 meV. This is well conceivable that the few-meV energy difference will be mitigated by a reasonable finite temperature (few tens of Kelvin), i.e., the spin-polarized states will transit into paramagnetic states in actual experiments. So that we do not consider the effect of spin polarization on the results for the present work.

In calculating the conductance for the junction under an external electric field, it is divided into the left electrode, the right electrode, and the middle scattering region. The SS matrix can be obtained by matching the wavefunctions at the two interfaces of electrode/scattering-region. Once the SS matrix is obtained, the conductance of the system at zero temperature can be calculated by using the Landauer formula Datta

GLR=e2hTLR=e2hnL,mR|Smn|2,G_{LR}=\frac{e^{2}}{h}T_{LR}=\frac{e^{2}}{h}\sum_{n\in L,m\in R}|S_{mn}|^{2}, (3)

where L/R labels the left/right leads, TLRT_{LR} is the transmission coefficient from lead L to lead R, and SmnS_{mn} gives the scattering amplitude from an incoming mode nn to an outgoing mode mm, both of which are the elements of the scattering matrix. It is clear that the conductance depends on the number of available transport modes through the junction.

III Results and Discussions

III.1 edge bands under in-plane electric field

Refer to caption
Figure 2: The band structure for (a-c) (1,3)PNR with NN=35 and (d-f) (2,4)PNR with NN=49 under an in-plane electric field with different strengths, (a and d) εy\varepsilon_{y}=0, (b and e) 2 mV/Å and (c and f) 5 mV/Å, respectively. The Fermi energy EFE_{F} is set to zero, and the left and right inserts in (b), (c) and (f) indicate the LDOS distributions corresponding to energies near EFE_{F} marked by the red and blue points, respectively, where the side view for each supercell LDOS is embodied. The right-side attached color bar indicates the electron density from the lowest (blue) to the highest (red) value.

The calculated energy spectra for the selected exemplary (1,3)PNR (NN=35) and (2,4)PNR (NN=49) with nearly the same width \sim3 nm under an in-plane electric field with different strength are shown in Figs. 2(a-c) and 2(d-f), respectively, where the Fermi level EFE_{F} is set to zero. The left and right inserts in (b, c and f) indicate the LDOS corresponding to energies marked respectively by the red and blue points contributed by the ribbon upper or lower edge outermost atoms, where the side view for each supercell is embodied. First, as shown in Fig. 2(a), the band structure of the pristine (1,3)PNR has a shuttle-shaped edge band near EFE_{F}, which implies the ribbon is metallic. In fact, the shuttle-shaped edge bands are twofold-degenerate differentiating by the red dashed and blue solid lines. However, the degeneracy can be eliminated by applying an in-plane electric field along the yy-direction. As a small field with strength εy\varepsilon_{y}=2 mV/Å is applied shown in Fig. 2(b), the twofold-degenerated edge bands obviously separate into two partially overlapped shuttles. Interestingly, as shown in Fig. 2(c), the two shuttles are completely separated as the filed strength is further increased up to 5 mV/Å, which means the degeneracy is completely removed. In contrast, for the (2,4)PNR as the example of the another category, it has two near-flat degenerate edge state bands in the original band structure which are drawn by the red dashed and blue solid lines shown in Fig. 2(d). They also pass through EFE_{F} exhibiting metallicity, which is very similar to that of a conventional zPNR except for a little difference in the curvature of the edge bands Carvalho ; Ezawa ; Guo . However, much different from the twofold-degenerated bands, the degeneracy for this category of ribbon can be easily removed by a small electric field as shown in Figs. 2(e) and 2(f). The characteristic of edge band under an electric field for (2,4)PNR is very similar to the extensively studied zPNR Ezawa ; ZhouBL ; Zhang , and hence we attribute them to the same category.

An in-plane electric field removing the degeneracy of the edge bands can be understood from Hamiltonian (1). The additional diagonal terms increase (accumulate) linearly with the yy-coordinate. From the analysis of the calculated results, it can be seen that the degenerate edge state bands for both categories of sPNRs would be broken by the application of an in-plane electric field. As the field strength increases, a band gap is opened and the transition from metal to semiconductor is occurred. The variation of (2,4)PNR edge bands with electric field is basically the same as zPNR Ezawa ; ZhouBL ; Zhang . This is because that they belong to the same category of sPNRs, where the outermost atoms of the ribbon are all at the same sublayer. However, for the (1,3)PNR with twofold-degenerate edge bands, the EFE_{F} is embedded in a mirrored shuttle-shape bands. In order to further know which atoms in the ribbon contribute the edge state bands, we have calculated the LDOS at the energies near EFE_{F} marked by the red and blue points, respectively. When the two shuttle-shape bands are not completely separated shown in Fig. 2(b) for εy\varepsilon_{y}=2 mV/Å, from the inserts we see the difference in LDOS between energies on the two shuttles. The red/blue dashed/solid shuttle bands are mainly contributed the outermost atoms at the upper/lower edge.

In addition, the left and right inserts in Figs. 2(c) and 2(f) depict the real-space electronic distributions at the energies of conduction band minimum (CBM) and valence band maximum (VBM) marked by the red and blue points, respectively. The right-side attached color bar indicates the density from the lowest (blue) to highest (red) value. From the LDOS inserts we can identify whether an edge band is completely contributed by the outmost atoms of the upper or lower edge of the ribbon, which provides a (upper-lower) edge DOF. However, we cannot distinguish them from which sublayers since their responses to an in-plane field are the same. Therefore, we need to apply an off-plane vertical electric field along the zz-direction respectively to the two categories of sPNRs.

III.2 edge bands under off-plane electric field

Refer to caption
Figure 3: The band structures for (a-d) (1,3)PNR with odd (NN=35) and even (NN=36) parities under an off-plane vertical electric field with different strengths, respectively, where (a and c) εz\varepsilon_{z}=0.1 V/Å and (b and d) 0.5 V/Å. The band structures for odd-even (NN=49-50) (2,4)PNR under the same field are correspondingly shown in (e-h). The inserts in (b, d, f and h) depict the LDOS distributions corresponding to the edge states at the certain energies pointed by the red and blue points, where the side view for each supercell LDOS is embodied. The color bar indicating the density is neglected here since it is the same as that in Fig. 2.

The difference in the layer atomic arrangement on the edge of a sPNR with different parity rows can be reflected by the response to an off-plane electric field, and then affects the corresponding band structure. This effect has been revealed and discussed for other 2D material nanoribbons, such as graphene, silicene and phosphorene ones ZhouBL ; LiZ ; Kang . Here, in Fig. 3 we present the result of the edge band response to a vertical electric field with strengths εz\varepsilon_{z}=0.1 and 0.4 V/Å for the two exemplary ribbons, respectively. For the consideration of the even-odd parity LiuP ; ZhouBL , we consider the two width cases of NN=35, 36 and NN=49, 50 for (1,3)PNR and (2,4)PNR, respectively. The LDOS left/right insert corresponding to the energies marked by the red/blue point. From Fig. 3(a) for (1,3)PNR with NN=35, the twofold-degenerate edge bands in shuttle-shape are almost unchanged under small εz\varepsilon_{z}=0.1 V/Å. However, when εz\varepsilon_{z} is increased up to 0.4 V/Å, interestingly, the two shuttle-shape edge bands are separated into two nearly flat bands as shown in Fig. 3(b). This results implying a transition from metal to semiconductor phase. But the two-fold degenerate is still not removed. Further, from Figs. 3(a-d) it seems that the number of atom rows (parity) does not affect the band structure of (1,3)PNR. In addition, the side view in (b) shows that the left/right inserted LDOS come from the edge outermost atoms of the top/bottom sublayer. We can identify a band from the top or bottom sublayer, which provides a sublayer DOF. This is because that the edge outmost atoms are alternately located at the top and bottom sublayers. In the contrast, for (2,4)PNR belongs to another category with two near-degenerate edge bands, the response of the edge bands to the field is sensitive on the ribbon width (parity), which is the same as that for the conventional zPNRs ZhouBL . As shown in Figs. 3(e) and 3(f), in specification, with the increase of the field strength the edge band degeneracy of the odd-numbered (NN=49) ribbon can not be broken but enterally move upward a little. Since the degenerate edge bands are contributed by the outermost atoms of both edges. However, for the even-numbered (NN=50) ribbon shown in Figs. 3(g) and 3(h), the outermost atoms of the two edges come from the same sublayer. Hence as the field strength increasing, the edge band degeneracy is removed and the transition from metal to semiconductor occurs.

III.3 (1,3)PNR homogenous junction

Next, from the above results for the two exemplary sPNRs belonging to different categories, we know that a monolayer (1,3)PNR owns two DOFs of edge and sublayer, which is particularly similar to a bilayer zPNR with two DOFs of edge and layer Soleimanikahnoj . Using this similarity, we may construct a homogenous pp-nn junction using a (1,3)PNR. The two sides of a (1,3)PNR are attached by the near top and bottom back-gates which adjust the EFE_{F}, leading to the lifting-up or -down of the edge bands for two ends. Further, the external in-plane and off-plane electric fields applied on the junction shown in Fig. 1(c) can be realized by using a side-gate and another back-gate electrodes, respectively. The side-gate technique has been proven to be experimentally feasible in graphene as a channel material for other applications Hahnlein ; Molitor . The back-gates can arouse a potential difference (electric field) across the whole monolayer. This kind of setup has been realized for bulk phosphorene transistors Tayari ; Kim . And a pseudospin field effect transistor has also been proposed and characterized based on a bilayer zPNR-based junction, in which a pseudospin-polarized current is generated Soleimanikahnoj . The similar interesting effect may also be realized in the pp-nn junction based on a monolayer (1,3)PNR, which can generate an edge- or sublayer-polarized current by properly adjusting the gate electrodes.

Refer to caption
Figure 4: The conductance spectrum for the (1,3)PNR-based pp-nn junction with a width about 3 nm under (a) in-plane transverse and (b) off-plane vertical electric fields, respectively. The gray solid line implies the conductance without applied field. The blue and red solid (dashed) lines indicate the edge bands from the same (different) (a) upper-lower edge with field strength 0.5 and 0.5 mV/Å, (b) top-bottom sublayer with field strength 0.1 and 0.5 V/Å are aligned, respectively. (c) and (d) The alignment of the edge bands of the two semi-infinitive ribbons under transverse and vertical fields, respectively, where the green/red arrow implies the on/off state of the junction.

In Fig. 4(a), we show the calculated conductance spectrum for the proposed junction based on (1,3)PNR with a width \sim3 nm under a transverse in-plane field with different strengths εy\varepsilon_{y}=0 (gray line), 0.5 (blue) and 5 mV/Å (red), respectively. The solid/dashed line indicates the alignment for the edge bands from the same/different upper-lower edge. The separated edge bands from the two edges of the terminals with/without alignment by back-gates is shown in Fig. 4(c), where the upper/lower panel corresponds to the case of the solid/dashed line in 4(a), and the green/red arrows implies the switch on/off state. First, we find a conductance plateau of 2e2/h2e^{2}/h around EFE_{F} for the junction without external field shown by the grey solid line, which is in accord with the band structure for a pristine (1,3)PNR shown in Fig. 2(a). When a small field 0.5 mV/Å is applied, by adjusting the positive and negative voltages of the two back-gate as to the edge bands from the same upper or lower edge be aligned, the conductance still maintains a 2e2/h2e^{2}/h plateau (blue solid line) near EFE_{F}, whereas the conductance is decreased to 0.3e2/h0.3e^{2}/h from different edge as shown by the blue dot-dashed line. This is because that the band degeneracy is broken not enough so that the two shuttle-shaped edge bands are not completely separated. Further, as the field strength increases up to 5 mV/Å, the degenerate shuttle-shaped edge bands is fully removed and a significant band gap is opened around EFE_{F} [see Fig. 2(c)]. As a consequence, when the same edge bands are aligned as shown in the upper panel in 4(c), a conductance gap correspondingly appears with two 2e2/h2e^{2}/h plateaus beside it (red solid line). Otherwise the conductance is almost zero even though there are two shuttle-shaped edge bands from the different upper or lower edge are aligned as shown in the lower panel in 4(c), which shows a transport off state.

Since the speciality of the edge atom arrangement for the monolayer (1,3)PNR, sublayer is also an important DOF in its transport property. In Fig. 4(b), we show the conductance spectrum for the junction under an off-plane vertical electric field with strengths, 0 (gray), 0.1 (blue) and 0.5 V/Å (red), respectively. Here the solid/dashed lines indicates the two edge bands from the same/different sublayers are aligned, of which the diagrammatic sketch for the edge band alignment is shown in 4(d). First, we also find a conductance plateau of 2e2/h2e^{2}/h around EFE_{F} for the junction without external field shown by the grey solid line. In the presence of applied field εz\varepsilon_{z}=0.1 V/Å, there is a gap opened within the shuttle-shaped edge bands. Therefore, as the edge bands from the same upper or lower edge are adjusted (by back-gate) to be aligned, the conductance exhibits two steps of 2e2/h2e^{2}/h besides EFE_{F} (see the blue solid line), which is in accord with the band structure shown in Fig. 3(b) or 3(d). In this case the increase of the field strength (e.g., εz\varepsilon_{z}=0.5 V/Å) only result in the two narrower steps and more far away from EFE_{F} (see the res solid line). On the contrary, when the edge bands from the same upper or lower edge are not aligned, the conductance is nearly zero except for very small peaks at EFE_{F} as shown by the blue and red dashed lines. The reason for the results is that the edge outermost atoms of (1,3)PNR are alternately located at the top and bottom sublayers. Therefore, a small vertical electric field can results in wave function superposition in different sublayers.

Refer to caption
Figure 5: The wavefunction overlap, <<ΨL(n,E)|ΨR(n,E)\Psi_{L}^{*}(n,E)|\Psi_{R}(n^{\prime},E)>>, as a function of (a) εy\varepsilon_{y} and (b) εz\varepsilon_{z} for the two terminals of the junction, where nn (nn^{\prime}) is the band index number for the junction left (right) side. And the LDOS diagrams with certain energies at two different field strengths pointed by black arrows are inserted with the embodied side view of supercell.

In order to further understand the rule of the two DOFs in transport for the junction, in Fig. 5 we plot the dependence of wavefunction overlap, <<ΨL(n,E)ΨR(n,E)\Psi^{*}_{L}(n,E)\Psi_{R}(n^{\prime},E)>>, as a function of the field strength (a) εy\varepsilon_{y} and (b) εz\varepsilon_{z}. Here ΨL/R(n,E)\Psi_{L/R}(n,E) is the wavefunction in the left/right (L/R) side of the junction. The LDOS with certain energies at two different field strengths are also inserted with the embodied side view of supercell. From Fig. 5, we first note that the overlap value drops rapidly as the field strength increasing due to the degeneracy elimination, with which the edge bands from the different upper-lower edge or top-bottom sublayer are separated. Therefore, without applied field the wavefunction overlap is equal to 1. In this case the edge bands are completely superposed, which refers to its pristine state shown in Fig. 2(a). Meanwhile, by adjusting the two back-gates as to the two side edge bands from the same edge or sublayer aligned as shown in upper panels of Fig. 4(c) or 4(d), the overlap degree is also equal to 1 regardless of the electric field strength.

Further, as shown in Fig. 5(a), we find that the two shuttle-shaped edge bands from the upper or lower edge are aligned by adjusting the back-gates as to the potential difference between the junction two sides is equal to the electric field strength [see the lower panel in Fig. 4(c)], the wavefunction overlap for two terminals is greatly reduced. When the two shuttle-shape edge bands are just separated a little by a small field 0.5 mV/Å, the wavefunction overlap is greatly reduced to a small value, which corresponds to the blue dashed peaks near EFE_{F} in Fig. 4(a). This is because that the overall degeneracy of the two shuttle bands would be almost removed even under a small electric field, but some of the degeneracy is still remained. Therefore, at certain energies the superposition of the wavefunctions in two sides still has a definitive value. In this case, from the LDOS shown in left upper panel of Fig. 5(a)], we can see it distributes at the upper and lower edge outermost atoms. Specifically, the upper edge atoms in the left side account for the major contribution (bright red spots), while the reddish ones at lower edge indicate a partial contribution. On the contrast, the LDOS distributed are the upper-lower edge for the right side is opposite to the left one. As the field strength is increased up to 5 mV/Å, the wavefunctions overlap for two side approaches zero (as red dashed line shown in Fig. 4(a)). This is because that the two shuttle edge bands are completely separated, and electrons are localized only at the outmost atoms of one edge. This can also be seen from the LDOS inserts with the embodied side view of supercell. On the other hand, since the edge mostout atoms of (1,3)PNR are alternately arranged on the top and bottom sublayers, the wavefunctions of the top and bottom sublayer atoms still mix together when a large vertical electric field is applied as shown in the lower insert of Fig. 5(b). Hence the amplitude scale of yy-axis in Fig. 5(b) is much larger than that in 5(a). Therefore, when εz\varepsilon_{z}=0.1 V/Å the wavefunction overlap for two sides is still large, which corresponds to the blue dashed plateau near EFE_{F} in Fig. 4(b). When the vertical field becomes larger, the wavefunction overlap of the top and bottom sublayers tends to zero due to the localization of electrons, as shown by the lower insert on Fig. 5(b) for LDOS with 0.5 V/Å.

Refer to caption
Figure 6: The conductance spectrum of a (1,3)PNR with monatomic vacancy respectively located at the (a, c) center and (b, d) edge of the ribbon, where (a, c) with εy\varepsilon_{y}=5 mV/Å and (c, d) εz\varepsilon_{z}=0.1 V/Å. The red dashed line in (a) and (c) corresponds to the pristine case without defect.

Finally, we demonstrate the influence of defect on the transport, which is also very important for device application by using the proposed DOFS since defects are inevitable on PNR samples in experiments Kiraly . For the sake of argument on the two DOFs in a (1,3)PNR, here we mainly consider two kinds of monatomic vacancies. As shown in the top of Fig. 6 for a defective (1,3)PNR, there are two monatomic vacancy positions indicated by the black dotted circles, one at the edge (EV) and the other in the center (CV) of the ribbon. Figure 6 gives the conductance spectrum, where (a, c) and (b, d) for CV and EV cases with εy\varepsilon_{y}=5 mV/Å and εz\varepsilon_{z}=0.1 V/Å, respectively. As a reference, the red dashed lines in (a, c) indicate the conductances for a pristine (1,3)PNR without defect but with in- and off-plane fields, respectively, which can be refer to the corresponding energy spectrum shown in Figs. 2 and 3. As is seen from the conductance spectra shown by the blue solid lines, a CV does not affect the edge bands regardless the field direction, resulting in an unchanged conductance in comparison the pristine case shown in (a, c). But the situation for EV is much different. Comparing with the conductance spectrum shown by the red solid line in Fig. 4(a) for the same upper-lower edge are aligned (also shown by red dashed line in Fig. 6(a)). As shown in Fig. 6(b), we find that the edge state on one of edges with an EV is severely weakened, which result in that the conductance plateau on the right side of EFE_{F} almost disappears. The original conductance plateau is maintained. In addition, for a EV with εz\varepsilon_{z}=0.1 V/Å shown by the blue solid line in Fig. 6(d), the conductance steps are retained as the ideal case [see the blue solid line in Fig. 4(b) and the red dashed line in Fig. 6(c)]. But the conductance peak on the left side of EFE_{F} is reduced by half, while the right side peak is still close to 2e2/h2e^{2}/h. This importantly implies that the CV has no effect on the edge state. The same result is expected for the in-plane field case. Moreover, due to the particularity of the two DOFs, the effect of EV on edge-state contributed conductance is different under the regulation of electric field. As shown by the LDOS in Fig. 5(a), a very small in-plane field can make the outer electrons of the edge atoms localized, which provides the transport channel corresponding to the edge with an EV almost forbidden. On the other hand, a large off-plane field would cause a certain degree of wavefunction overlap between the sublayer atoms on the same side [as shown in Fig. 5(b)], so that an EV in one sublayer only destroys the edge state of the corresponding sublayer, while the other sublayer on the same side still maintains a channel. In general, the sPNRs with two-DOFs can reduce the influence of low-concentration edge monatomic defects on the edge states to a certain extent.

At last but not least, besides the atomic defects the edge passivation of nanoribbons is also an important issue because that the edges are usually passivated in real experimental samples. When edge hydrogen(H)-passivation is considered in the TB calculation, usually an extra potential field is applied on the edge atoms of the ribbons. This may lead to that the original edge bands are disappeared or modified (see, e.g., LiuP ). Here we have also tested the H- and O-passivation cases by using a DFT calculation, respectively. The result shows that the H-passivated edges do not exhibit edge states, while O-passivated ones remain qualitatively similar to the the case of bare edges. And the result of the edge bands is the qualitatively the same as those by the TB approach, which confirms that the device assumption based on (1,3)PNR is scientifically valuable Ren .

IV Summary and Conclusion

In summary, we have studied the difference in electronic structures between the two categories of sPNRs with edge states and the variation under an external electrostatic field. Taking (1,3)PNR and (2,4)PNR as the examples, we first identify the distinction of edge morphology and the unit cell selection for these two ribbons. And then, by using the KWANT code based on TB approach, we find that the shuttle-shaped twofold-degenerate edge bands of (1,3)PNR can be became two separated shuttles until the degeneracy is removed and a gap near EFE_{F} is opened under a sufficient strong in-plane electric field. And each shuttle contributed from the outmost atoms of the ribbon upper or lower edge is verified according to the LDOS. However, under a small off-plane field the shuttle-shape bands are easily separated into two degenerated near-flat bands contributed from the edge atoms of the top or bottom sublayer. The edge band variation with external field for this category is completely different from that of (2,4)PNR belonging to previously reported zPNR category. This is because a (1,3)PNR has four zigzag atomic configurations on the upper and lower edges and the degenerate bands are from the outermost atoms in the same sublayer or different upper-lower edge. This allows the two DOFs to be distinguished and regulated by applying electric field along different directions. Further, based on this issue we propose a (1,3)PNR homogenous junction attached by electric gates. Interestingly, the transport property of the junction with field manipulation well reflects the characteristics of the two DOFs for (1,3)PNR category. In addition, the defect effect from the vacancies in edge and bulk on the transport property is also discussed. These results may provide a further understanding on PNRs and initiate new developments in PNR-based electronics.

Acknowledgements.
This work was supported by the National Natural Science Foundation of China (Grant Nos. 11774085, 11804092, and 11664019), the Project Funded by China Postdoctoral Science Foundation (Grant Nos. BX20180097, 2019M652777), and Hunan Provincial Natural Science Foundation of China (Grant No. 2019JJ40187).

References

  • (1) K. S. Novoselov, A. K. Geim, S. V. Morozov, D. Jiang, Y. Zhang, S. V. Dubonos, I. V. Grigorieva, and A. A. Firsov, Science 306, 666 (2004).
  • (2) K. F. Mak, C. Lee, J. Hone, J. Shan, and T. F. Heinz, Phys. Rev. Lett. 105, 136805 (2010).
  • (3) P. San-Jose, E. Prada, E. McCann, and H. Schomerus, Phys. Rev. Lett. 102, 247204 (2009).
  • (4) X. Xu, W. Yao, D. Xiao, and T. F. Heinz, Nat. Phys. 10, 343 (2014).
  • (5) S. Soleimanikahnoj, and I. Knezevic, Phys. Rev. Appl. 8, 064021 (2017).
  • (6) Yi Ren, Pu Liu, Benliang Zhou, Xiaoying Zhou, and Guanghui Zhou, Phys. Rev. Appl. 12, 064025 (2019).
  • (7) Likai Li, Yijun Yu, Guo Jun Ye, Qingqin Ge, Xuedong Ou, Hua Wu, Donglai Feng, Xian Hui Chen, and Yuanbo Zhang, Nat. Nanotech. 9, 372 (2014).
  • (8) Xi Ling, Han Wang, Shengxi Huang, Fengnian Xia, and Mildred S. Dresselhaus, PNAS 112, 4523 (2015).
  • (9) C. Q. Han, M. Y. Yao, X. X. Bai, Lin Miao, Fengfeng Zhu, D. D. Guan, Shun Wang, C. L. Gao, Canhua Liu, Dong Qian, Y. Liu, and Jin-feng Jia, Phys. Rev. B 90, 085101 (2014).
  • (10) Likai Li, Jonghwan Kim, Chenhao Jin, Guojun Ye, Diana Y. Qiu, Felipe H. da Jornada, Zhiwen Shi, Long Chen, Zuocheng Zhang, Fangyuan Yang, Kenji Watanabe, Takashi Taniguchi, Wencai Ren, Steven G. Louie, Xianhui Chen, Yuanbo Zhang, and Feng Wang, Nat. Nanotech. 12, 21 (2017).
  • (11) Steven P. Koenig, Rostislav A. Doganov, Hennrik Schmidt, A. H. Castro Neto, and Barbaros Özyilmaz, Appl. Phys. Lett. 104, 103106 (2014).
  • (12) Michele Buscema, Dirk J. Groenendijk, Sofya I. Blanter, Gary A. Steele, Herre S. J. van der Zant, and Andres Castellanos-Gomez, Nano Lett. 14, 3347 (2014).
  • (13) Fengnian Xia, Han Wang, and Yichen Jia, Nat. Commun. 5, 4458 (2014).
  • (14) Sherman Jun Rong Tan, Ibrahim Abdelwahab, Leiqiang Chu, Sock Mui Poh, Yanpeng Liu, Jiong Lu, Wei Chen, and Kian Ping Loh, Adv. Mater. 30, 1704619 (2018).
  • (15) A. Carvalho, A. S. Rodin, and A. H. Castro Neto, Europhys. Lett. 108, 47005 (2014).
  • (16) Han Liu, Adam T. Neal, Zhen Zhu, Zhe Luo, Xianfan Xu, David Tomanek, and Peide D. Ye, ACS Nano 8, 4033 (2014).
  • (17) Pengke Li and Ian Appelbaum, Phys. Rev. B 90, 115439 (2014).
  • (18) Xiaoying Zhou, Wen-Kai Lou, Dong Zhang, Fang Cheng, Guanghui Zhou, and Kai Chang, Phys. Rev. B 95, 045408 (2017).
  • (19) X. Y. Zhou, R. Zhang, J. P. Sun, Y. L. Zou, D. Zhang, W. K. Lou, F. Cheng, G. H. Zhou, F. Zhai, and Kai Chang, Sci. Rep. 5, 12295 (2015).
  • (20) Pu Liu, Xianzhe Zhu, Xiaoying Zhou, Guanghui Zhou, and Kai Chang, Sci. China-Phys. Mech. Astron. 64, 217811 (2021).
  • (21) Paul Masih Das, Gopinath Danda, Andrew Cupo, William M. Parkin, Liangbo Liang, Neerav Kharche, Xi Ling, Shengxi Huang, Mildred S. Dresselhaus, Vincent Meunier, and Marija Drndic, ACS Nano 10, 5687, (2016).
  • (22) Mitchell C. Watts, Loren Picco, Freddie S. Russell-Pavier, Patrick L. Cullen1, thomas S. Miller, Szymon P. Bartuś, Oliver D. Payton, Neal T. Skipper, Vasiliki tileli, and Christopher A. Howard, Nature 568, 216 (2019).
  • (23) Vy Tran and Li Yang, Phys. Rev. B 89, 245407 (2014).
  • (24) M. Ezawa, New J. Phys. 16, 115004 (2014).
  • (25) Rui Zhang, X. Y. Zhou, D. Zhang, W. K. Lou, F Zhai, and Kai Chang, 2D Mater. 2, 045012 (2015).
  • (26) Ajanta Maity, Akansha Singh, Prasenjit Sen, Aniruddha Kibey, Anjali Kshirsagar, and Dilip G. Kanhere, Phys. Rev. B 94, 075422 (2016).
  • (27) Marko M. Grujić, Motohiko Ezawa, Milan Ž. Tadić, and Francois M. Peeters, Phys. Rev. B 93, 245413 (2016)
  • (28) Benliang Zhou, Benhu Zhou, Xiaoying Zhou, and Guanghui Zhou, J. Phys. D: Appl. Phys. 50, 045106 (2017).
  • (29) K. Nakada, M. Fujita, G. Dresselhaus, and M. S. Dresselhaus, Phys. Rev. B 54, 17954 (1996).
  • (30) M. V. Bollinger, J. V. Lauritsen, K. W. Jacobsen, J. K. Nørskov, S. Helveg, and F. Besenbacher, Phys. Rev. Lett. 87, 196803 (2001).
  • (31) D. J. P. de Sousa, L. V. de Castro, D. R. da Costa, and J. M. Pereira, Jr. Phys. Rev. B 94, 235415 (2016).
  • (32) SKFiroz Islam, Paramita Dutta, A. M. Jayannavar, and Arijit Saha, Phys. Rev. B 97, 235424 (2018).
  • (33) Yi Ren, Fang Cheng, Z. H. Zhang, and Guanghui Zhou, Sci. Rep. 8, 2932 (2018).
  • (34) E. Taghizadeh Sisakht, F. Fazileh, M. H. Zare, M. Zarenia, and F. M. Peeters, Phys. Rev. B 94, 085417 (2016).
  • (35) Ashwin Ramasubramaniam and Andre R. Muniz, Phys. Rev. B 90, 085424 (2014).
  • (36) Yuanyue Liu, Fangbo Xu, Ziang Zhang, Evgeni S. Penev, and Boris I. Yakobson, Nano Lett. 14, 6782 (2014).
  • (37) Yangjin Lee, Sol Lee, Jun-Yeong Yoon, Jinwoo Cheon, Hu Young Jeong, and Kwanpyo Kim, Nano Lett. 20, 559 (2020).
  • (38) Liangbo Liang, Jun Wang, Wenzhi Lin, Bobby G. Sumpter, Vincent Meunier, and Minghu Pan, Nano Lett. 14, 6400 (2014).
  • (39) B. Kiraly, N. Hauptmann, A. N. Rudenko, M. I. Katsnelson, and A. A. Khajetoorians, Nano Lett. 17, 3607 (2017).
  • (40) L. L. Li, and F. M. Peeters, Phys. Rev. B 97, 075414 (2018).
  • (41) C. W. Groth, M. Wimmer, A. R. Akhmerov, and X. Waintal, New J. Phys. 16, 063065 (2014).
  • (42) Yongjin Jiang and Liangbin Hu, Phys. Rev. B 75, 195343 (2007).
  • (43) M. Zwierzycki, P. A. Khomyakov, A. A. Starikov, K. Xia, M. Talanana, P. X. Xu, V. M. Karpan, I. Marushchenko, I. Turek, G. E. W. Bauer, G. Brocks, and P. J. Kelly, Phys. Stat. Sol. B 245, 623 (2008).
  • (44) Tatiane P. Santos, Leandro R. F. Lima, Caio H. Lewenkopf, J. Comput. Phys. 394, 440 (2019).
  • (45) A. N. Rudenko and M. I. Katsnelson, Phys. Rev. B 89, 201408(R) (2014).
  • (46) S. Smidstrup, T. Markussen, P. Vancraeyveld, J. Wellendorff, J. Schneider, T. Gunst, B. Verstichel, D. Stradi, P. A. Khomyakov, and U. G. Vej-Hansen, J. Phys.: Condens. Matter 32, 015901 (2020).
  • (47) S. Datta, Electronic transport in mesoscopic systems (Cambridge University Press, Cambridge, 1995).
  • (48) Hongyan Guo, Ning Lu, Jun Dai, Xiaojun Wu, and Xiao Cheng Zeng, J. Phys. Chem. C, 118, 14051 (2014).
  • (49) Dong Zhang, Wenkai Lou, Maosheng Miao, Shou-cheng Zhang, and Kai Chang, Phys. Rev. Lett. 111, 156402 (2013).
  • (50) Z. Li, H. Qian, J. Wu, B. L. Gu, and W. Duan, Phys. Rev. Lett. 100, 206802 (2008).
  • (51) Jun Kang, Fengmin Wu, and Jingbo Li, Appl. Phys. Lett. 100, 233122 (2012).
  • (52) B. Hähnlein, B. Hündel, J. Pezoldt, H. Töpfer, R. Granzner, and F. Schwierz, Appl. Phys. Lett. 101, 093504 (2012).
  • (53) F. Molitor, J. Guttinger, C. Stampfer, D. Graf, T. Ihn, and K. Ensslin, Phys. Rev. B 76, 245426 (2007).
  • (54) V. Tayari, N. Hemsworth, O. Cyr-Choinière, W. Dickerson, G. Gervais, and T. Szkopek, Phys. Rev. Appl. 5, 064004 (2016).
  • (55) J. S. Kim, P. J. Jeon, J. Lee, K. Choi, H. S. Lee, Y. Cho, Y. T. Lee, D. K. Hwang, and S. Im, Nano Lett. 15, 5778 (2015).