Brought to you by:
Paper

The electronic transport efficiency of a graphene charge carrier guider and an Aharanov–Bohm interferometer

, and

Published 13 November 2018 © 2018 IOP Publishing Ltd
, , Citation Xuan Wei et al 2018 J. Phys.: Condens. Matter 30 485302 DOI 10.1088/1361-648X/aae9d3

0953-8984/30/48/485302

Abstract

The electrostatic gating defined channel in graphene forms a charge carrier guider. We theoretically investigated electronic transport properties of a single channel and an Aharanov–Bohm (AB) interferometer, based on a charge carrier guider in a graphene nanoribbon. Quantized conductance is found in a single channel, and the guider shows high efficiency in the optical fiber regime, in good agreement with the experiment results. For an AB interferometer without a magnetic field, quantized conductance occurs when there are a few modes inside the channel. The local density of states (LDOS) inside the AB interferometer shows quantum scars when the scattering is strong. At low magnetic field, a periodical conductance oscillation appears. The conductance has a maximum value at zero magnetic field in the absence of intravalley scattering. The mechanism was investigated by LDOS calculations and a toy model.

Export citation and abstract BibTeX RIS

1. Introduction

Graphene is a potential candidate for the next generation of electronic devices because of its peculiar electronic, mechanical and chemical properties [13]. Graphene nanoribbons, which act as a hard-wall confinement of electrons, were usually fabricated as a quasi-one-dimensional (1D) electron conducting material [48]. Ribbons usually have defects at the edges, therefore scattering is inevitable. In the absent of a strong magnetic field, the linear conductance is usually smaller than a quantized value [912]. With newly developed techniques, it is possible to fabricate an atomically precise narrow graphene nanoribbon with arbitrary width or chirality [1318]. Alternatively, to eliminate scattering, an electrostatic gating defined quasi-1D charge carrier channel with smooth boundaries can be formed by gating graphene flakes [1928]. The smooth boundaries form a soft-wall confinement channel. Electrons inside the channel behave in a same way that light travels inside an optical fiber, which is an interesting aspect of Dirac electrons as an analogy to optics [1922]. The Fermi level of the Dirac electrons, with respect to the charge neutral point, has the same role as the refractive index in optics. The channel has high transport efficiency when: (i) in the optical fiber (OF) regime, i.e. the charge carrier density inside the channel is larger than outside; (ii) in the bipolar regime, i.e. the charge carriers inside and outside the channel are electron-like and hole-like (or vice versa), respectively [2426].

The gate defined charge carrier guider was studied in a recent experiment [26]. The observed conductance is (4n  +  2)e2/h with n an integer and e2/h the quantized conductance. It was speculated that valley preserved transmission happens inside the guider. In previous theoretical works [29, 30], the transport properties of a gate defined charge carrier guider device were simulated using a tight-binding model. It was found that the conductance shows a plateau at (4n  +  2)e2/h when the charge carrier density outside the channel is small, in agreement with the experiment [26]. The conductance 4ne2/h results from both spin and valley degeneracy, and the extra 2e2/h comes from the edge states (either the zigzag edged or zigzag-armchair mixed edged) of the graphene nanoribbons [3134]. Nevertheless, in another experimental work, it was proposed that the conductance through the channel should be 4ne2/h in a suspended graphene with gate defined channel [25], considering the non-uniform of the gated graphene region and limited transmission probability. The different conductance values reported in the above two references are due to different configuration of the devices. Kim et al [26] reported a two-terminal device, and the injected electrons, after leaking out the channel, are involved in the transport through the edge states. Rickhaus et al [25] reported a four-terminal device, and the leaking electrons tend to transmit into two lateral terminals. Therefore, in such a device the edge states will not contribute to the whole transmission.

In graphene Aharanov–Bohm (AB) interferometers, due to their long mean-free paths [35, 36], the quantum interference of massless Dirac fermions were also investigated both in theory [3744] and experiment [26, 4554]. In experiments, etched graphene AB interferometers are mainly used for transport measurements; AB interferometers defined by fluorination [52] and local gate were investigated as well [26]. In the local gate defined AB interferometer, the influenced of edge roughness is weak due to smooth boundaries, therefore, the transport of edge states and the boundary scattering is suppressed. It was found that the density of states inside the AB ring forms quantum scars, which should be attributed to disorders [52, 53]. In all the above mentioned experimental works, the conductance varies periodically when the magnetic flux (or magnetic field) changes during a period of h/e (the quantized magnetic flux). According to the Onsager relation, the conductance through an AB interferometer should have either a maximum or a minimum at the zero magnetic field. In [26] and [51], the conductance has minimal values at zero magnetic field. While in [45], the conductance has a maximal value at zero magnetic field. In [48] and [54], both maximal and minimal values of conductance were observed under different gate voltages. One may wonder how the interference at zero magnetic field is determined in specific experiments.

In this paper, we theoretically investigate the electronic transport properties of a four-terminal gate defined charge carrier channel and an AB interferometer in graphene (figure 1(a)). The conductance in a single straight channel shows plateau values (4ne2/h) when the gate voltage changes, which means both a well function of the charge carrier guider with weak scattering, and the absence of edge states transmission. The transmission efficiency is large in the OF regime. The results are in good agreement with the experimental results [25, 26]. For an interferometer, the conductance shows clear plateaus when there are only a few propagating states. The local density of states (LDOS) inside the channel forms quantum scars when the scattering is strong, which can be verified by the scanning gate microscopy techniques [52, 53, 55]. Under a weak magnetic field, the conductance oscillates with the magnetic flux period h/e. When there are only two modes (valley degenerate states) in the channel, the interference shows a maximal value at zero magnetic field. As the number of the modes inside the channel increases, either a maximum or a minimum of the conductance appears. We attribute the results to intravalley scattering [56], which is explained by a toy model.

Figure 1.

Figure 1. (a) The schematic of a four-terminal monolayer graphene charge carrier guider. In the central region, the gate voltage in the light cyan region is Vin and is Vout otherwise, thus a channel is formed. The blue solid line (red dotted line) represents the trajectory of charge carriers within (beyond) the OF regime. The width and length of the graphene are N  =  18 and L  =  27 (the number of hexagonal units). The width of the channel is M  =  10. The transmission coefficients T13 (b), T12 (c) and T24 (d) and η (e) versus Vin are shown in two cases: $ V_{\rm out}=0.005$ and $V_{\rm out}=-0.005$ . The black line in (b) is the transmission coefficient for an ideal zigzag edged graphene nanoribbon with the same width as the channel. The other parameters in (b)–(e) are N  =  200, M  =  40 and L  =  200.

Standard image High-resolution image

The rest of the paper is organized as follows: in section 2 the Hamiltonian and the calculation methods are specified. Numerical results are displayed in section 3, including the properties of multi-terminal transmission coefficients of a single straight channel and an AB interferometers under weak magnetic field. LDOS in both cases was also calculated to facilitate the discussion concerning the number of propagating states and chaotic transport related quantum scars. A toy model is adopted to explain the quantum interference of AB interferometer around the zero magnetic field and the role of intravalley scattering. Finally, a brief conclusion is displayed in section 4.

2. Model and methold

The device we considered is displayed in figure 1(a), a graphene nanoribbon (the central region) with four-terminal connected. To form a charge carrier guider in the central region, larger values were set for the charge carrier density inside the channel than outside, which can be realized by tuning the gate voltages. The Hamiltonian in the tight-binding representation is written as [57, 58]:

Equation (1)

The first term is the on-site potential which is determined by the gate voltage. In figure 1(a), the gate voltage in the light cyan region was set to $ \newcommand{\e}{{\rm e}} \epsilon_i=V_{\rm in}$ , and $ \newcommand{\e}{{\rm e}} \epsilon_i=V_{\rm out}$ in the white region. Therefore, when $|V_{\rm in}|>|V_{\rm out}|$ (the OF regime), a graphene charge carrier guider is formed in the middle. The on-site potential of terminal 1 and 3 (2 and 4) are uniformly determined by gate voltage Vin (Vout). The second term describes the coupling between the nearest sites with energy t. For a perpendicular magnetic field, a phase $\phi_{ij}$ induced by the vector potential $\bf A$ is added in the hopping term: $\phi_{ij}=\frac{2\pi e}{h}\int_{i}^{\,j} {\bf A}\cdot {\rm d}{\bf l}$ . Under a gauge transformation, $\phi_{ij}=m\phi$ was set for the mth row atoms along the lateral direction. So the actual magnetic induction strength B is related to ϕ by $B=\oint{\bf A}\cdot {\rm d}{\bf l}/S=\frac{h\phi}{2\pi eS}$ with S the area of a single hexagonal unit.

The spin degenerate current flow through a four-terminal graphene nanoribbon, which was calculated from the Landauer–Büttiker formula, is $I_{ij}=\frac{2e}{h}\int T_{ij}(E) {\rm d}E$ [59]. Here the transmission coefficient Tij(E) from terminal i to j was calculated using the Green's function method [58]:

Equation (2)

with $\boldsymbol{\Gamma}_{i}$ the line-width function of terminal i and ${\bf G}^{\rm r}/{\bf G}^{\rm a}$ the retarded/advanced Green's function of the central region. ${\bf G}^{\rm r}$ was calculated from the relation $ \newcommand{\re}{{\rm Re}} \renewcommand{\dag}{\dagger}{\bf G}^{\rm r}=[{\bf G}^{\rm a}]^\dagger=(E{\bf I}-{\bf H}_{\rm c}-\sum_i\boldsymbol{\Sigma}_{i}^{\rm r}){}^{-1}$ with ${\bf H}_{\rm c}$ the Hamiltonian of the central region and $\boldsymbol{\Gamma}_{i}^{\rm r}$ the retarded self-energy of the ith terminal. Finally $\boldsymbol{\sum}_{i}^{\rm r}$ can be calculated numerically either by solving eigen-equations or iterations [60, 61].

By setting the voltage in terminal 2, 3 and 4 to zero, the current flow from terminal 1 to other terminals (driven by a tiny voltage bias $\Delta V$ at zero temperature) is proportional to the transmission coefficient in equation (2). To directly compared to the experimental results, the linear conductance was simply set as $G=\frac{2e^2}{h}T(E)$ . Therefore, in the following discussion, only the transmission coefficients are discussed and they represent the magnitude of the linear conductance and charge current flow. In the numerical calculation, we used t in equation (1) as the energy unit and set the Fermi level E  =  0.

3. Numerical results

In a real experiment, the potential at the channel boundaries cannot be changed in a step way. So in the numerical calculation, a relaxation region was introduced by varying $ \newcommand{\e}{{\rm e}} \epsilon_i$ in equation (1) between Vout and Vin smoothly in a length of five hexagonal units. In a previous report, it was shown that the states inside the channel are immune to the profile of the potential variation [30].

3.1. Straight charge carrier guiders

We start by discussing the characteristic of a single gate defined charge carrier guider, as shown in figure 1(a). In figure 1(b), transmission coefficient T13 versus Vin is shown for $V_{\rm out}=0.005$ and $V_{\rm out}=-0.005$ . The transmission coefficient of an ideal graphene nanoribbon (the same width as the channel) is also displayed for comparison. For an ideal graphene nanoribbon, T13 is quantized with value 2n  +  1. For the charge carrier guider, T13 shows plateaus of value 2n. As $|V_{\rm in}|$ increases, a new plateau indicates two valley extra degenerate modes are propagating inside the channel [25]. The positions of the plateau edges are highlighted by the vertical dotted lines. Compared with the curve for an ideal graphene nanoribbon, the transmission coefficient of a charge carrier guider has a difference of 1. It means that the transmission of the edge states through the charge carrier guider is eliminated: the edge states flow from terminal 1 into two lateral terminals (2 and 4). In previous studies [26, 29, 30], the conductance in a two-terminal graphene guider is (2n  +  1)e2/h. The edge states transmit along the boundary of the ribbon contribute e2/h to the total conductance. In our work, the present model is a four-terminal device, similar to [25], thus the conductance is only contributed by the states inside the channel. When comparing these two curves, it can be noticed that the conductance for $V_{\rm out}=0.005/V_{\rm out}=-0.005$ is large when Vin is positive/negative, respectively. For example, for positive Vin, T13 for negative Vout is better quantized with flat plateaus. The density of states at the channel boundary is depleted, when the gate voltage across the boundaries changes its sign, and the modes are well restricted inside the channel.

Now we focus on the transmission coefficients from terminal 1 to terminal 2 and 4. As T14 and T12 are the same considering device symmetry, we only display the dependence of T12 on Vin. In figure 1(c), when $|V_{\rm in}|>0.02$ , T12 remains around 0.4. Therefore, the total transmission coefficient from terminal 1 to terminal 2 and 4 is close to unity. Several small peaks (guided by the grey dotted line) can be noticed. The peak positions are the same as the new appeared conductance plateaus in figure 1(b). It is contributed by an arising mode leaking from the channel. If we assume that a single channel is confined in the y direction and the charge carriers flow in the x direction (hence kx is a good quantum number), then, for a given Fermi level E, the relationship between the two momentum components can be expressed as $E=3at\sqrt{k_x^2+k_y^2}/2$ , with a the nearest C–C bond length in graphene and t the nearest coupling energy in equation (1). When an emerging mode appears in the channel, a low momentum kx results in a small injecting angle to the channel boundaries (see the red dotted line in figure 1(a)). So the mode is likely to leak out the channel through the Klein tunneling [62, 63]. As the Fermi energy increases, the kx is increased and the incident angle is large. When the incident angle becomes larger than the critical angle, the mode propagates well in the channel (see the blue solid line in figure 1(a)). From the above discussion, it is not surprising to see that T24 versus Vin show a few peaks. These peaks are contributed by the resonant tunneling of the charge carriers across the channel, in which a Fabry–Pérot interferometer is formed [64, 65]. The peaks become higher when Vin and Vout have opposite signs, because the confinement of the states in the channel becomes stronger. To study the transmitting efficiency of the charge carrier guider, we defined $ \newcommand{\e}{{\rm e}} \eta=T_{13}/(T_{13}+T_{12}+T_{14})$ and the results are shown in figure 1(e). When $|V_{\rm in}|$ is small, η is small as well. As $|V_{\rm in}|$ increases, η increases and shows several plateaus. The results agree with the experiment work in [25].

To further demonstrate transport of confined states inside the channel, the spatial LDOS $\rho_{\bf i}$ of a two-terminal graphene nanoribbon are shown for different Vin in figure 2. Here $\rho_{\bf i}$ was calculated from the Green's function in the central region $\rho_{\bf i}(V_{\rm in})=-\mathrm{Im}[{\bf G}_{\bf i}^{\rm r}(V_{\rm in})]/\pi$ [66]. For small values of Vin (e.g. $V_{\rm in}=0.1$ in figure 2(a)), there are two valley degenerate states (corresponding to T13  =  2 in figure 1(b)), and the LDOS inside the channel shows a single peak. For slightly large values as $V_{\rm in}=0.135$ and 0.15 in figures 2(b) and (c), the LDOS inside the channel has two peaks, which correspond to two states at each valley, and the four states make a transmission contribution of 4. As Vin further increases (figures 2(d)(f)), more states are propagating inside the channel and the number of LDOS peaks are related to the transmission coefficient (divided by 2). The multi-peaks are evenly distributed in the transverse direction. It is worth noting that, for $V_{\rm in}=0.135$ and $V_{\rm in}=0.21$ (figures 2(b) and (d)), the LDOS expands beyond the channel boundaries, indicating current leaking. At these two Vin values, new T13 plateaus only occurs with small kx, and newly appeared states tend to leak out the channel. Interestingly, in all cases in figure 2, the LDOS shows periodical quantum scars along the channel. The quantum scars in figures 2(b), (d) and (f) are stronger than the others, which can be explained by the diffusive transport of confined relativistic Dirac fermions in graphene [67, 68].

Figure 2.

Figure 2. LDOS of graphene guider channels for different Vin values. The sample sizes are the same to figures 1(b)(e) and $V_{\rm out}=-0.005$ .

Standard image High-resolution image

We now discuss the dependence of Vin and Vout on transmission coefficients. Figure 3 shows two-dimensional (2D) maps of T13, T12, T24 and η as functions of Vin and Vout. In figure 3(a), T13 has small values near $V_{\rm in}=0$ , indicating an unformed charge carrier guider. As Vin increases, plateau appears for T13. At small $|V_{\rm out}|$ values, the plateaus become clear when Vin and Vout have opposite signs (the bipolar regime). At large $|V_{\rm out}|$ values, the plateaus become indistinct. Larger T13 values can be found in the region where Vin and Vout have the same sign (the OF regime), because the current flow outside the channel can contribute. T12 and T24 show similar features: the values are small when either Vin or Vout is close to zero, and the values are large when both Vin and Vout are large. In addition, oscillation pattern was found for T24 when $V_{\rm out}>0$ , $V_{\rm in}<0$ or $V_{\rm out}<0$ , $V_{\rm in}>0$ (the bipolar regime), because of the resonant tunneling across the charge carrier channel. In figure 3(d), η has high guiding efficiency when $|V_{\rm in}|>|V_{\rm out}|$ (the OF regime). When $|V_{\rm in}|<|V_{\rm out}|$ , the guiding efficiency η is small. Similar results are found in a recent experimental work [25].

Figure 3.

Figure 3. T13 (a), T12 (b), T24 (c) and η (d) as functions of Vin and Vout for the device in figure 1(a). The sample sizes are the same to figures 1(b)(e).

Standard image High-resolution image

3.2. Aharonov–Bohm effect in a graphene ring

Next we investigate the electronic transport properties in a ring composed of two charge carrier guiders, as shown in figure 4(a). The relationships between T13 and Vin are shown in figure 4(b) for both two-terminal and four-terminal devices. Both curves show vibration and step behaviors at low $|V_{\rm in}|$ , indicating that there are only a few modes in the channel. When $|V_{\rm in}|$ is large, T13 shows strong irregular vibrations, which can be attributed to the intravalley scattering. We will discuss it later in details. The Vin values between two neighbor plateaus are the same to figure 1(a). T13 values for the four-terminal device are smaller than two-terminal device, which means that the edge states leak out of the channel to lateral terminals (2 and 4).

Figure 4.

Figure 4. (a) The schematic of a four-terminal graphene AB interferometer defined by charge carrier guiders. The channel region is shown in light cyan. The transmission coefficients T13 for a four-terminal device and a two-terminal device (terminal 2 and 4 are disconnected) are shown in (b). The inset in (b) displays the schematic of a two-terminal AB interferometer device. The other parameters in (b) are N  =  400, M  =  40, L  =  400 and $V_{\rm out}=-0.005$ .

Standard image High-resolution image

The LDOS of two-terminal devices are shown in figure 5 for different Vin values, guided by arrows in figure 4(b). For $V_{\rm in}=0.04$ , the AB ring carries no state and the transmission is mainly through the ribbon edges (figure 5(a)). In figures 5(b)(h), the states are well restricted within the channel. Two main features can be found: (i) the quantum scars occurs inside the channel. Periodical LDOS pattern is more obvious (figures 5(b) and (e)) when the propagating states are just formed. In such a case, kx is smaller and the scattering is strong (figure 4(b)). In experiment, the radial quantum scars were observed in the ring area [52, 53]. The appearance of quantum scars is associated to the chaotic nature of states in a cavity, in which the conductance fluctuates. (ii) When there are more than two states in each arm of the AB ring (e.g. in figures 5(e)(g)/(h), there are four/six states), scattering happens between these states. Since the states are valley degenerate, the scattering is an intravalley one [56]. Under a weak magnetic field, the scattering changes the dynamic phase of each state, and then the interference varies accordingly.

Figure 5.

Figure 5. (a)–(h) The LDOS of an AB interferometer in two-terminal graphene devices for different Vin when $\phi=0$ . (i)–(l) The LDOS of AB rings in different shapes: (i) a narrower AB ring, (j) a diamond shaped AB ring, (k) a half AB ring (i.e. a curve channel) and (l) an incomplete AB ring with a gap in the bottom arm. In (i)–(l), the gated region is illustrated by the LDOS maps and $V_{\rm in}=0.15$ . The ribbon sizes (N and L) are the same to figure 4.

Standard image High-resolution image

To demonstrate how the quantum scars are affected by the shape of the AB ring, four kinds of AB rings are shown in figures 5(i)(l) with $V_{\rm in}=0.15$ and $V_{\rm out}=-0.005$ . In figure 5(i), LDOS pattern of a narrow ring is similar to that in figure 5(e). For a diamond shaped AB ring shown in figure 5(j), LDOS has a zigzag pattern. For a half AB ring (i.e. a curve single channel) where the confinement effect is weak, the quantum scars disappear, and the transport of charge carriers is nearly ballistic. For an incomplete ring with a gap (figure 5(l)), the quantum scars are obvious, and the states are scattered strongly at the gap, and the incomplete ring serves as a cavity. In experiments, such gate voltage dependent chaotic transport and quantum scars are expected using scanning gate microscopy techniques [52, 53, 55].

In figure 6, 2D maps of T13 (versus ϕ and Vin) are displayed for both two-terminal and a four-terminal devices. The magnetic field is rather weak, therefore quantum Hall edge states are not formed, and the AB effect is predominant. In both figures, when Vin is smaller than 0.13, clear AB interference can be observed as ϕ changes. There are two valley degenerate propagating modes inside the channel. The period of the oscillation is about h/e, indicating the well performance of the interferometer. T13 shows a maximum at $\phi=0$ . When $V_{\rm in}>0.13$ , irregular oscillations appear in addition to the regular oscillation with period h/e. Figure 7 gives a more clear demonstration of the oscillations. For example, when $V_{\rm in}=0.22$ and 0.23, T13 shows a maximum (figures 7(a) and (b)). While for $V_{\rm in}=0.25$ and 0.28, T13 shows a minimum (figures 7(c) and (d)). There are more than two modes propagating inside the channel. The minimal value at zero magnetic field should be attributed to the intravalley scattering (figures 5(f) and (h)). In previous experimental work, the linear conductance of an AB ring at zero magnetic field either shows a maximum or a minimum [26, 45, 48, 50, 54]. Here we show that this behavior is due to scattering between discrete propagating states inside a single valley, i.e. intravalley scattering.

Figure 6.

Figure 6. Aharonov–Bohm interference in graphene charge carrier guider AB interferometer. T13 are displayed for ϕ and Vin is displayed in two conditions: (a) a two-terminal device (figure 4(b) inset) and (b) a four-terminal device (figure 4(a)). (c)–(e) Simple AB rings models for the explanation of the interference process in three conditions: a single propagating mode (c), two independent propagating modes (d) and two propagating modes with scattering considered (e).

Standard image High-resolution image
Figure 7.

Figure 7. T13 are displayed for ϕ relations for different Vin (along the red dashed lines in figure 6(b)) in a four-terminal AB interferometer: Vin = 0.22 (a), Vin = 0.23 (b), Vin = 0.25 (c) and Vin = 0.28 (d).

Standard image High-resolution image

Now we explain the above results with a toy model. Figures 6(c)(e) display symmetric AB rings sandwiched by two leads in three conditions: (i) a single mode, (ii) two modes with no scattering, and (iii) two modes with scattering. The incoming wave from the left lead, which is split at the left cross section of the interferometer, after traveling along both arms of the interferometer, flows into the right lead. The interference is determined both by the magnetic flux and the dynamic phase acquired by propagating waves.

  • (i)  
    When there is only a single mode in the channel (figure 6(c)), the total transmission coefficient contributed by two waves from the upper and lower arms is given by $\tau=|{\rm e}^{{\rm i}(kl+\Phi/2)}+{\rm e}^{{\rm i}(kl-\Phi/2)}|={2+2\cos \Phi}$ with k the momentum along the channel, l the length of each arm and Φ the magnetic phase enclosed by the interferometer. So τ always shows a maximum at $\Phi=0$ .
  • (ii)  
    When there are two independent modes in the channel (figure 6(d)). The transmission coefficient can be expressed as:
    Equation (3)
    with $\delta k=k_1-k_2$ . So for two independent modes, there is always a constructive interference and τ is a maximum when $\Phi=0$ .
  • (iii)  
    When there are two modes in the channel (figure 6(d)), under scattering, the two states may interchange (figure 5(f)). For the simplest case, we assume that the charge carrier of momentum k1 in the upper arm, after traveling a distance l1, is scattered to the momentum k2 and then propagates a distance of l2 ($l_2=l-l_1$ ), to the right lead. Meanwhile, the charge carriers of momentum k2 are transferred to the momentum k1 due to the same process. The transmission coefficient reads:
    Equation (4)

Easy to notice that only the last term in the equation (4) is Φ dependent. If $\delta k=0$ , τ always shows a maximum when $\Phi=0$ . If $\cos\frac{\delta kl}{2}\cos\frac{\delta k (l_1-l_2)}{2}<0$ , τ shows a minimum when $\Phi=0$ .

The above discussion explains our numerical results. When there are only two modes (valley degenerated) in the channel, T13 are displayed for ϕ shows periodical oscillation and a constructive interference at zero magnetic field, because the intervalley scattering is rather small [29]. When there are more than two modes (e.g. four, six or more modes), T13 are displayed for ϕ shows irregular oscillation pattern due to the scattering. Since the propagating modes for a single valley are quite nearby to each other in the momentum space, the scattering is caused by the intravalley scattering. In all cases, T13 is either a maximum or a minimum at $\phi=0$ .

In the experiments of AB interferometers, when the conductance is larger than 2e2/h, both constructive interference and destructive interference can occur due to the scattering. When the conductance is smaller than 2e2/h, there are two possibilities. First, there are more than two modes and the conductance is suppressed to less than 2e2/h by either disorder or scattering. Second, there are only two modes and the disorder or scattering is rather weak due to a clean graphene. Only for the latter case the uniform constructive interference can be realized. Thus the quantum interference and the role of intravalley scattering in graphene AB interferometer can only be verified in controllable atomically precise samples in which the scattering can be deliberated to be on or off. With the present state-of-the-art graphene techniques, high-quality AB interferometers can be fabricated (e.g. with bottom–up approach) in which either the width or the charge carrier population can be precisely controlled by doping or a gate voltage [13, 16, 18]. So our finding for quantum interference in graphene AB ring around the zero magnetic field can be confirmed by experiment with nowaday graphene techniques. The spatial resolved states and chaotic transport related quantum scars in AB interferometer can also be detected by using scanning gate microscopy techniques [52, 53, 55].

4. Discussion and conclusion

In this work, the electronic transport characteristics of a single electrostatic gating defined charge carrier guider and an AB interferometer composed of two charge carrier guiders were therotically studied. For a single charge carrier guider, the conductance shows plateau values of 4ne2/h. The efficiency of carrier guider is high in the OF regime. The results are in good agreement with the experimental work of [25] and [26]. The spatial distribution and the number of propagating states inside the channel can be seen from the LDOS maps. For an AB interferometer, the conductance shows a periodical oscillation in a period of h/e under varying the magnetic field. When there are only two modes inside the channel, the conductance shows maximum at the zero magnetic field. When there are more propagating modes, the conductance shows extra irregular oscillations and the conductance can be either a minimum or a maximum around zero magnetic field due to the intravalley scattering. Quantum scars appear in the LDOS map when the scattering is strong and the intravalley scattering can also be seen from the LDOS. A simple model is applied to describe the role of intravalley scattering in the interference of AB rings.

Acknowledgments

S G Cheng thanks R Z Zhang for careful reading of the manuscript. This work was supported by NSFC under grant Nos. 11674264 and 11874298. Wen-Jing Zhang is financially supported by the Scientific Research Program Funded by Shaanxi Provincial Education Department No. 17JK0786.

Please wait… references are loading.
10.1088/1361-648X/aae9d3