Quantum transport of Dirac fermions in selected graphene nanosystems away from the charge-neutrality point
Abstract
Peculiar electronic properties of graphene, including the universal dc conductivity and the pseudodiffusive shot noise, are usually attributed to a small vicinity of the charge-neutrality point, away from which electron’s effective mass raises, and nanostructures in graphene start to behave similarly to familiar Sharvin contacts in semiconducting heterostructures hosting two-dimensional electron gas. Using the effective Dirac equation for low-energy excitations it can be shown that, as long as abrupt potential steps separate the sample area from the leads, some graphene-specific features can be identified even relatively far from the charge-neutrality point. Namely, the conductance is reduced, by a factor varying from to (depending on the sample geometry) comparing to the standard Sharvin value, whereas the shot noise is amplified, with the Fano factor varying from to . In this paper, we confront the results of earlier analytic considerations with numerical simulations of quantum transport on the honeycomb lattice, for selected systems containing up to lattice sites, with different geometries for which considerations starting from the Dirac equation cannot be directly adapted. The results show that for a wedge-shape constriction with zigzag edges and approximately square shape of the narrowest section, the transport characteristics can be tuned from graphene-specific sub-Sharvin values to standard Sharvin values, depending on whether the electrostatic potential profile in the narrowest section is rectangular (i.e., has abrupt steps) or smooth (approximately parabolic). Similarly, the half-Corbino disk with zigzag edges and rectangular potential profile exhibits both the conductance and the noise close to the sub-Sharvin values. In contrast, for a circular quantum dot with two narrow openings and irregular edges, the conductance is close to the Sharvin value, and the Fano factor approaches the value of , characterizing the symmetric chaotic cavity. For a similar quantum dot with a hole, eliminating direct trajectories connecting the two openings, the conductance is reduced by a factor close to , coinciding with the sub-Sharvin value, but the Fano factor is almost unaffected remaining close to . This suggests that, in experimental attempt to verify the predictions for sub-Sharvin transport regime, one should focus rather on nanosystems with relatively wide sample openings, for which the scatterings on edges are insignificant next to the scatterings on sample-lead interfaces.
I Introduction
There are few phenomena in nature for which the results of measurements of physical quantities are given directly by the fundamental constants of nature, leaving even the question of the actual number of fundamental constants open Duf02 ; Leb19 . In the second half of the last century, two phenomena from this group were discovered and theoretically described: the quantum Hall effect And75 ; Kli80 ; Lau81 ; Tsu82 and the Josephson effect Jos74 , which are currently used as the basis for the standards of units of resistance and electric voltage in the SI system, i.e., the ohm Pay20 and the volt Tan15 . The discovery of the two-dimensional allotrope of carbon, graphene, made at the beginning of the 21st century Nov05 ; Zha05 allowed for the improvement of the Ohm standard based on the quantum Hall effect Pay20 . (Some peculiar features of the Josephson effect in graphene were also pointed out Tit06 ; Sal23 .) Moreover, it turned out that several material characteristics of graphene, such as the conductivity Kat06 ; Kat20 or visible light absorption Nai08 ; Sku10 , are given by the fundamental constants or dimensionless numerical coefficients. The sub-Poissonian shot noise (quantified by the Fano factor ) Two06 ; Dan08 ; Ryc09 ; Lai16 and the anomalous Lorentz number Yos15 ; Cro16 ; Ryc21a ; YTT23 for charge-neutral graphene also can be regarded as examples of such characteristics. Although measurements of these quantities with metrological accuracy are not possible yet, the scientific community have undoubtedly gained unique opportunities to test a theoretical model, which is the effective two-dimensional Dirac-Weyl equation for monolayer graphene Sem84 ; DiV84 .
The author and Witkowski have recently find, using the effective Dirac equation, that for doped graphene samples of highly-symmetric shapes (namely, the rectangle with smooth edges and the Corbino disk) the conductance is reduced, whereas the shot noise is amplified, comparing to standard Sharvin values Ryc21b ; Ryc22 . The reduction (or amplification) is maximal when abrupt potential steps separate the sample area from the leads; for instance, the conductance (with Sha65 ; Wee88 ; Gla88 , the conductance quantum, the Fermi wavenumber, and the sample width diskfoo ) for the rectangle or the narrow disk (i.e., the inner-to-outer radii ratio ), for the wide-disk limit (); the Fano factor for the rectangle or the disk with , for the disk with . When the potential profile gets smoothen the above-listed sub-Sharvin values evolve towards and the Fano factor approaches the ballistic value of . Later, the discussion in analytical terms was extended on the nonzero magnetic field case Ryc23 ; Ryc24 , showing that in a doped disk with the vanishing conductance (notice that in the disk geometry the edge states are absent and the current is blocked at sufficiently high field except from narrow resonances via Landau levels DaS09 ; Ryc10 ; Zen19 ; Sus20 ; Kam21 ) is accompanied by a non-trivial value of .

It is the purpose of this paper to extend the discussion of sub-Sharvin transport regime in graphene by going beyond the effective Dirac equation. In particular, we address the question how reallistic (irregular) edges of a nanosystem carved out of the honeycomb lattice affect transport characteritics? For this purpose, we perform computer simulations of quantum transport for selected systems depicted schematically in Fig. 1 modeled within the tight-binding Hamiltonian. The results show that sub-Sharvin characteristics are closely reconstracted for relatively short and wide systems; for longer and more complex system with multiple constrictions some less obvious scenarios (including the sub-Sharvin conductance accompanied by the shot-noise power resembling a chaotic cavity) can be observed.
The remaining parts of the paper are organized as follows. In Sec. II, we present the Landauer-Büttiker formalism for a generic nanoscopic system and key literature results following from the effective Dirac equation for graphene at the charge-neutrality point as well as in the sub-Sharvin regime. Higher charge-transfer cumulants for graphene at and away from the charge-neutrality point are also discussed in Sec. II. Statistical distributions of transmission probabilities for different quantum-transport regimes, including the sub-Sharvin transport regime in graphene, are described in Sec. III. The tight-binding model of graphene and our main results concerning the conductance and the Fano factor for selected nanosystems (see Fig. 1) are presented in Sec. IV. The conclusions are given in Sec. V.

II Landauer-Büttiker transport in nanoscopic systems and graphene
II.1 Remark on the origin of zero-temperature Landauer-Sharvin resistance
First, let’s look for a concise answer to the question: Where does electrical resistance come from at absolute zero?
In the familiar Drude model of electrical conduction Kit05 electrons are assumed to constantly bounce between heavier, stationary lattice ions, allowing one to express the material specific resistivity as a function of the electron’s effective mass, velocity, and the mean free path. In quantum-mechanical decription of solids, Drude model provides a reasonable approximation as long as the Fermi wavelength remains much shorter than electron’s mean free path and the conductor size.
The picture sketched above changes substantially when electic charge flows through a nanoscopic system, such as quantum point contact in semiconducting heterostructure Wee88 , a carbon nanotube Whi98 , or monoatomic quantum wire Agr03 (see Fig. 2, top part). Assuming for simplicity that such a system has no internal degrees of freedom leading to the degeneracy of quantum states, in other words — that in a sufficiently small energy range we have at most one quantum state (level) — we note that the time of flight of an electron through the system is limited from below by the time-energy uncertainty relation
(1) |
Next, by linking the energy range with the electrochemical potential difference in macroscopic electrodes (reservoirs) connected to the nanoscopic system (see Fig. 2, bottom part), we can write
(2) |
where denotes the difference in electrostatic potential on both sides of the system, and is the elementary charge (without sign). Combining the above equations we obtain the limit for the electric current flowing through the system,
(3) |
which means that the electrical conductivity
(4) |
We thus see that the uncertainty principle of energy and time leads to a finite value of the conductivity, and therefore to a nonzero value of the electrical resistance, of a nanoscopic system. By rigorous derivation, the upper bound in Eq. (4) is replaced by , introducing the Landauer-Sharvin resistance in noninteracting electron systems Sha65 ; Lan57 ; But85 . Obviously, many-body effects may alter this conclussion substantially. For instance, the resistivity of graphene sample may drop below the Landauer-Sharvin bound due to hydrodynamic effects Kum22 . In twisted bilayer graphene, both the interaction-driven insulating and superconducting (i.e., resistance-free) phases were observed Cao18a ; Cao18b ; Fid18 . These issues are, however, beyond the scope of the present wotk.
II.2 The Landauer-Büttiker formula
At a temperature close to absolute zero () and in the limit of linear response, i.e., the situation in which the electrochemical potential difference also tends to zero (), it can be shown that the electrical conductivity of a nanoscopic system is proportional to the sum of transition probabilities for the so-called normal modes in the leads Naz09 ,
(5) |
where denotes the conductance quantum; namely, for systems exhibiting spin degeneracy (for graphene, we have due to the additional degeneracy — called valley degeneracy — related to the presence of two nonequivalent Dirac points in the dispersion relation). The probabilities () are calculated by solving (exactly or approximately) the corresponding wave equation (Schrödinger or Dirac) for a fixed energy, which, given the assumptions made, can be identified with the Fermi energy . Importantly, we perform the calculations under the additional assumption that there are so-called waveguides between the macroscopic reservoirs and the nanoscopic system, for which we can provide (for a fixed value of ) solutions in the form of propagating waves, the number of which is or , for the left or right waveguide, respectively (see Fig. 2). We also assume that a wave that reaches the waveguide-reservoir interface is never reflected. It is worth noting that the sum appearing in formula (5) is the trace of the transmission matrix, the value of which does not depend on the choice of the basis; therefore, it can be expected that the result does not depend on how precisely we construct the aforementioned waveguides, which, it is worth emphasising, are an auxiliary construction that usually has no direct physical interpretation. (For the same reason, the result will be the same whether we consider scattering from left to right or in the opposite direction.)
As mentioned above, the details of the calculations (or computer simulations) leading to the determination of the probability values () will depend on the geometry of the system under consideration. If waveguides are modelled as strips of fixed width (), at the edges of which the wave function disappears, the normal modes have the form of plane waves planewfoo , for which the longitudinal component of the wave vector () is continuous and the normal component () is quantized according to the rule
(6) |
The calculations are particularly simple in cases where the central region (marked with a dark square in Fig. 2) differs from the leads only in that it contains an electrostatic potential that depends on the coordinate (oriented along the main axis of the system), for example in the form of a rectangular barrier. Then the transmission matrix has a diagonal form (no scattering between normal modes occurs), and in special cases, such as the rectangular barrier mentioned above, but also e.g. the parabolic potential considered by Kemble in 1935 Kem35 , it is possible to provide compact analytical formulas.
We will not present the exact results here, but only point out that for solutions obtained by the mode-matching method for the Schrödinger equation, one can write approximately
(7) |
which we write more briefly as , with denoting the Heaviside step function. In Eq. (7) we introduce the wave vector corresponding to the Fermi energy (assuming for simplicity that the dispersion relation is isotropic) calculated with respect to the top of the potential barrier in the central region. Furthermore, assuming that there are many modes for which (which occurs if ), and therefore the summation in Eq. (5) can be replaced to a good approximation by integration, we obtain — via Eqs. (6) and (7) — the result known in the literature as the Sharvin conductance Sha65
(8) |
It is worth noting that the reasoning leading to Eq. (8) can be relatively easily applied to the case where the electrostatic potential in the central region is approximately constant and the width of the conducting region is a function of the position along the longitudinal axis (), changing slowly enough that the scattering between normal modes can be neglected. The above-mentioned case is the so-called quantum point contact (QPC), shown schematically in Fig. 2 (top part), which can be realized in semiconductor heterostructures hosting a two-dimensional electron gas (2DEG) Naz09 .
II.3 Shot noise and counting statistics
The second quantity, besides electrical conductivity, that characterizes nanoscopic systems at temperatures close to absolute zero is the shot-noise power. For the sake of brevity, let us point out the basic facts: First, the electric charge flowing through the system shown schematically in Fig. 2 (lower part) in a short time interval is a random variable. Second, the expectation value of such a variable is closely related to the electrical conductivity in the linear-response limit,
(9) |
The reason the measured value of fluctuates at successive time intervals is due to the discrete (granular) nature of the electric charge.
Assuming (for the moment) that electrons jump from one reservoir to another completely independently, we conclude that the charge flow is a Poisson process, or more precisely, that the quantity follows the Poisson distribution; the variance is therefore proportional to the expectation value given by Eq. (9),
(10) |
More generally, -th central moment can be written as
(11) |
with the integer .
The Fano factor, quantifying the shot-noise power, is defined as the ratio of the actual measured variance of the charge flowing through the system to the variance given by Eq. (10), or more precisely
(12) |
(For compact derivation, see e.g. Ref. Naz09 .) In the following, we have limited our considerations to long time intervals such that ; hence characterizes the zero-frequency noise, not to be confused with the celebrated noise in electronic systems Tak93 . A generalization of Eq. (12) for finite times (and nonzero temperatures) is also possible Sch07 .
In particular, it follows from Eq. (12) that the Poisson limit () is realized in the case of a tunnel junction, for which we have for each . This is a completely different case than the ballistic system considered above, which exhibits Sharvin conductance; then, replacing the summation with integration as before and using the approximation given by Eq. (7), we obtain
(13) |
In general, for fermionic systems we always have ; the factor appearing in the numerator in Eq. (12) is a consequence of the Pauli exclusion principle. In the case of the idealized ballistic system we have , see Eq. (13), which means that the electron count () does not fluctuate with time. One could say that the electrons avoid each other so much that they "march" at equal intervals. (Of course, this is only possible at absolute zero temperature, otherwise additional thermal noise appears, i.e. the Nyquist-Johnson noise proportional to the conductivity value, whose influence we have ignored here; see Ref. Naz09 .)
In an attempt to determine higher charge cumulants, it is convenient to introduce characteristic function
(14) |
such that
(15) |
Assuming for simplicity, we arrive at the Levitov-Lesovik formula Naz09 ; Sch07
(16) |
expressing the full counting for noninteracting fermions.
Substitution of the above into Eq. (15) with and reproduces (respectively) Eqs. (5) and (12). Analogously, for and , we get
(17) | ||||
(18) |
For the Sharvin regime, see Eq. (7),
(19) |

II.4 Scattering of Dirac fermions in two dimensions
Using the introductory information gathered above, we will now calculate — with some additional simplifying assumptions — the electrical conductivity as well as the higher charge cimulants of a graphene strip. The effective wave equation for itinerant electrons in this two-dimensional crystal is the Dirac-Weyl equation, the detailed derivation of which can be found, e.g., in Katsnelson’s textbook Kat20 , and which can be written in the form
(20) |
where the energy-independent Fermi velocity is given by , where denotes the nearest-neighbor hopping integral in the graphene plane and is the lattice constant (as a result, the approximate value of is about , which is several times lower than typical Fermi velocities in metals). The remaining symbols in Eq. (20) are the quantum mechanical momentum operator (the notation here means differentiation with respect to the selected coordinate, ), is a vector composed of Pauli matrices valleyfoo , and the electrostatic potential energy is assumed to depend only on the position along the principal axis of the system.
The above assumptions imply that we can look for solutions to Eq. (20) in the form of a two-component (i.e., spinor) wave function
(21) |
where and are functions of . By substituting the above ansatz into Eq. (20) we obtain a system of ordinary differential equations
(22) | ||||
(23) |
where the primes on the left-hand side denote derivatives with respect to . We see that in the system of Eqs. (22), (23) the quantities and play the role of parameters on which the solutions depend (in the following, when calculating, among others, the electrical conductivity, we will identify the electron’s energy with the Fermi energy by setting ).
At this point it is worth to comment on the problem of quantizing the value of the transverse momentum () in Eqs. (22), (23). Assuming that the component of the current density perpendicular to the axis of the graphene strip disappears at its edges (i.e., for and , see Fig. 3(a)), what is known as the so-called mass confinement Ber87 , we get a slightly different quantization than in the case of the Schrödinger system, see Eq. (6), namely
(24) |
In practice, however, the assumptions made in the following part mean that when calculating measurable quantities (, , etc.) we will approximate the sums appearing in Eqs. (5), (12), (17), (18), with integrals with respect to ; the quantization change described above is therefore insignificant for further considerations.
The solution of the system of Eqs. (22), (23) is particularly simple in the case if the electrostatic potential energy, i.e. the function , is piecewise constant. Then, the solutions in individual sections (i.e., areas where is constant) have the form of plane waves. For instance, for waves traveling in the positive and negative directions along the axis, look as follows
(25) |
where we have defined
(26) | ||||
For , propagating-wave solutions also exist (this is, by the way, the main difference between the solutions of the massless Dirac equation and the Schrödinger equation, which leads in particular to the phenomenon known as Klein tunneling Rob12 ; Gut16 ) and differ from those given in Eq. 25 only in some signs. We leave the straightforward derivation to the reader.
At the interface of regions differing in the (locally constant) value of , we perform a matching of wave functions, which for the two-dimensional Dirac equation reduces to solving the continuity conditions for both spinor components matchfoo . For instance, if we consider the scattering from the right side of the discontinuity to the left side, we write
(27) |
where the spinor functions with indices and differ in the values of and [see Eqs. (25), (26)], but are characterized by the same value of . Since the considerations concern the interface between the graphene sample and the graphene region covered with a metal electrode (see Fig. 3(a)), the calculations can be simplified by adopting the model of a heavily doped electrode, in which we set , where ; we can then write the wave functions on the left in asymptotic form
(28) |
where we have omitted the phase factor, which is not important for further considerations. After substituting the above into Eq. (27), the calculations are straightforward; we now present the results for the transition and reflection probabilities
(29) |
which turn out to depend only on the angle of incidence of the plane wave, or — more precisely — on the value of . In particular, we see that for we have (and ), which is a manifestation of the Klein tunneling mentioned above (let us emphasize that the potential barrier considered here has an infinite height).
The probability of passing through the entire graphene sample, i.e., through two electrostatic potential steps occurring at the sample-lead interface (see Fig. 3(a)), is most easily calculated using the double-barrier formula, the clear derivation of which can be found, e.g., in the Datta’s handbook Dat97
(30) |
where a phase shift
(31) |
related to the propagation of a plane wave along the main axis , is introduced. (Note here that the phase shift introduced in this manner also implies the assumption that any reflections from the side edges of the system do not change the value of ; in practice, this implies that we restrict our considerations to systems for which .) Assuming barrier symmetry, , , and substituting the formulas given in Eq. (29), we can now write explicitly as a function of and ,
(32) |
where
(33) |
and the Fermi wave vector, assuming for the sample region, is equal to . The absolute value in the last expression arises from the fact that formulas in Eq. (29) and the following results are identical for ; we leave the verification of this property to the reader.
II.5 The conductivity, shot noise, and higher cumulants for ballistic graphene strip
The physical consequences of the above expression for the transition probabilities , see Eqs. (32) and (33), are now discussed for two physical situations: a charge-neutral sample () and the Sharvin limit (). (Unless otherwise stated, we also assume geometry with long, parallel sample-lead interfaces and .)
In the first case () we obtain and can use the identity , resulting in a surprisingly simple expression
(34) |
visualized in Fig. 3(b). In the wide-sample limit, , the sums appearing in the formulas for Landauer conductance [see Eq. (5)], Fano factor [Eq. (12)], and higher cumulants , [Eqs. (17) and (18)] can be approximated with integrals [see also Eq. (24)], leading to
(35) | ||||
(36) |
or, more generally,
(37) |
where is the Euler gamma function. To facilitate future comparisons with other transport regimes, we will additionally define
(38) |
such that . Taking into account the graphene-specific fourfold degeneracy of states due to the presence of spin and valley degrees of freedom (the conductance quantum is therefore ), we obtain
(39) | |||
(40) |
The value of (instead of , as in a typical ballistic system) means that charge-neutral graphene exhibits universal specific conductivity, , the value of which is additionally determined only by the universal constants of nature. The value of the Fano factor is also not accidental, as it is a value characteristic for ohmic (disordered) conductors. (The same applies to higher cumulants.) In the context of graphene, the term pseudodiffusive conductivity is often used to emphasize that this ballistic system perfectly emulates an ohmic conductor within the appropriate parameter range. It should be emphasized that the first two theoretical values, given in Eq. (39) and originally derived in Refs. Kat06 (conductance) and Two06 (Fano factor), have been experimentally confirmed with satisfactory accuracy in 2008 Dan08 . (For the comprehensive theoretical discussion of full counting statistics for graphene at the Dirac point, see Ref. Sch10 .)


In the Sharvin limit () the situation looks a bit different. We can then assume that the contribution of modes for which (i.e., evanescent modes) is negligible and limit the considerations to . Next, we notice that as the values of (or their powers) are summed, the term of Eq. (32) oscillates rapidly, especially as approaches . Therefore, it seems reasonable to replace the sine argument with a random phase and average the result, which leads to the following approximation
(41) |
where we used the table integral Ryz-2-553-3
(42) |
substituting
(43) | ||||
The comparison between the approximation given in Eq. (41) and the actual , see Eqs. (32), (33), for , is presented in Fig. 3(d).
Eq. (41) is essentially the Dirac version of Eq. (7) describing a standard ballistic system; when calculating the Landauer conductance, we can again approximate the summation by integration and obtain
(44) |
where we recall the value of Sharvin conductance given in Eq. (8). The prefactor is a consequence of the fact that in the last expression in Eq. (41), where previously there was a step function , a term describing an arc of a circle has appeared; the conductance of graphene beyond the charge-neutrality point is therefore reduced compared to a typical ballistic system.
Interestingly, in deriving Eq. (44) we did not explicitly assume, as in Eq. (39), that the width to length ratio of the sample is ; hypothetically, the result given in Eq. (44) can therefore be applied whenever the condition is satisfied, regardless of the value of . In practice, however, it is difficult to imagine that the double-barrier transmission formula, Eq. (30), which is the basis of the entire reasoning, could be applied to samples that do not satisfy the condition. It seems that in the case of graphene samples with , hard-to-control edge effects can significantly alter the conductivity Mia07 . We will address this issue later in this paper, but first let us calculate the approximate values of the Fano factor and higher cumulants in the Sharvin limit.
In order to construct the approximation analogous to this in Eq. (41), but for , it is sufficient to calculate
(45) |
where we used the first derivative of , see Eqs. (42) and (43), with respect to . More generally, the -th power of can be approximated by
(46) | ||||
where is the hypergeometric function Abr65 . (Notice that for a positive integer , or is a non-positive integer, so the function reduces to a polynomial of ; after multiplying by , the resulting expression can be further simplified to a degree polynomial of the variable .) For and , the above reproduces the results of Eqs. (41) and (45), respectively. The next two expressions are
(47) | ||||
(48) |
Setting for , and approximating the summations occuring in Eqs. (12), (17), and (18) by integrations, namely,
(49) | ||||
(50) |
we obtain
(51) |
(For the first four numerical values of , see Table 1.)
The surprising (non-zero) value of the shot noise in Eq. (51) is close to the experimental results obtained by Danneau et. al. Dan08 , which are in the range of . (The aspect ratio of the sample used in this experiment was .) It should be emphasized that measuring the shot noise of nanoscopic devices containing components of different materials is rather challenging; there are also results in the literature that suggest that the dependence of the shot noise on the system filling is weak, with the value always close to the pseudodiffusive Dic08 . Measurements of higher charge cumulants are so far missing.
In Fig. 4, the approximations for , , , and , both for and for , are compared with and the actual values following from Eqs. (32), (33) for . Briefly speaking, the higher cumulant is considered, the larger value of is necessary to observe the convergence to the sub-Sharvin limit; however, for and for the heavily-doped leads, is sufficient. The discussion of more realistic situations (finite doping in the leads, other sample shapes) is presented later in this paper.
II.6 The narrow-opening limit
Using the conformal mapping technique Ryc09 ; Gui08 , it can be shown that for charge-neutral graphene (), the pseudodiffusive values given in Eqs. (37), (39), (40) are essentially valid for an arbitrary sample shape, provided that the prefector (if present) is replaced by an appropriate geometry-dependent factor defined by the conformal transformation. In particular, when mapping the rectangle onto the Corbino disk, one needs to substitute , where and are the outer and inner disk radii (respectively), see Fig. 5(a). An additional condition is that the system must be in the multimode regime, i.e. (or ) in the disk case. Otherwise, if , a nonstandard tunneling behavior is observed, with and Ryc09 . (Since the transport is governed by a single mode, we also have and in such a case.)
However, in the Sharvin limit ( in the disk case), a different set of universal charge-transport characteristics is predicted Ryc22 . Regardless of the exact size or shape of the outer sample-lead interface, one can assume that the double-barrier formula, Eq. (30), is still applicable and that and due to the Klein tunneling. Therefore, , the role of a phase shift is negligible, and one can write — for the wave leaving the inner lead with the total angular momentum (with ) — the transmission probability as
(52) |
Subsequently, the summation for the -th power of can be approximated (notice that we have assumed ) by integration over , leading to
(53) |
where, also using the symmetry , we have defined
(54) |
The first four values of are listed in Table 1. Substituting the above into Eqs. (5), (12), (17), and (18), we obtain
(55) | |||||
For the Corbino disk, it is also possible to perform the analytical mode matching for the heavily-doped leads, but arbitrary disk doping and radii ratio Ryc09 ; Ryc20 . We skip the details of the derivation here and just give the transmission probabilities
(56) |
where
(57) |
and [] is the Hankel function of the first [second] kind. In Fig. 5, we provide the comparison between the values of the conductance and the next three charge cumulants obtained by substituting the exact -s given above into Eqs. (5), (12), (17), (18), and performing the numerical summation over with the approximate values given in Eq. (55). It is easy to see that the radii ratio of is sufficient to reproduce our predictions for the narrow-opening limit, with good accuracy, typically starting from . (This time the higher cumulant is considered, the faster the convergence.)
Transport regime (or approximation) | ||||
Cumulant | Pseudodiffusive, | Sub-Sharvin, | , | Narrow-opening, |
see Eq. (WW) | ||||
Expressions for with arbirary are given by: a)Eq. (38); b)Eqs. (46), (50); c)Eq. (…); d)Eq. (54). | ||||
e)For , , with , coincides with . |
III Distributions of transmission probabilities
A compact and intuitive representation of the charge-transfer cumulants discussed in the previous Section is provided within the distribution function of transmission probabilities . This function takes the simplest form when the transmission probability can be expressed as a monotonic function of the parameter , i.e., . In such a case, the probability density is defined by
(58) |
where is the number of transmission channels per unit of [here constant and determined by the appropriate quantization rule, see Eqs. (6), (24), or (52)], and is the derivative of the inverse function . In a generic situation, the right-hand side in Eq. (58) needs to be replaced by the sum over the monotonicity intervals of .
In the pseudodiffusive limit, and , the transmission probability given by Eq. (34) immediately implies
(59) |
where we recall the pseudodiffisive conductance given in Eq. (39). The distribution is visualized in Fig. 3(c).
Analyzing the sub-Sharvin transport, we now change the order of presentation by switching to the disk geometry to point out that in the narrow-opening limit, i.e., for and , the transmission probability given by Eq. (52) leads to another closed-form expression for the distribution, namely
(60) |
with for a circular lead.
In the case of parallel interfaces at a distance , see Fig. 3(a), the description of the sub-Sharvin transport becomes more complex, since the transmission probability , see Eqs. (32) and (33), is no longer a monotonic function of . The distribution obtained numerically for is presented in Fig. 3(e), where the continuous corresponds to the limit. It can be noticed that each of the seven transmission minima [see Fig. 3(d)] produces a distinct (integrable) singularity of located at . A closed-form, asymptotic expression for in the limit is missing; instead, we propose the approximation directly following from given by Eq. (41), i.e.,
(61) |
Subsequent approximations for the cumulants can be evaluated as
(62) |
The numerical values for are listed in Table 1, together with the corresponding approximations for the charge-transfer cumulants , , and , which are also depicted in Figs. 4(b–d) [thick horizontal lines]. We notice that these values typically match the incoherent ones, obtained by substituting Eq. (46) into Eq. (50), within the accuracy that allows unambiguous identification of the transport regime. A surprising exception is the case of , for which the proximity of the pseudodiffusive () and incoherent () values is merely a coincidence. [By definition, the conductance .]
The functional forms of derived in this Section, along with a selection of others previously reported in the literature, can be found in Table 2.
Transport | System | Refs. | |||
---|---|---|---|---|---|
regime | |||||
Standard | Sharvin contact in 2DEG, | Wee88 ; Gla88 | |||
ballistic | width | ||||
(Pseudo)- | diffusive conductor | Bee92 ; Nag92 | |||
diffusive | |||||
Charge-neutral graphene sample | , | Two06 ; Dan08 | |||
(width , length ) | |||||
Charge-neutral graphene disk | Ryc09 | ||||
(inner radii , outer radii ) | |||||
Sub-Sharvin | Doped graphene sample | Ryc21b , | |||
(width , length ) | this work | ||||
Doped graphene disk, | Ryc22 , | ||||
(inner radii , outer radii ) | this work | ||||
Doped graphene disk, | — | Ryc24 | |||
Chaotic | Symmetric cavity | Bee97 | |||
a)Reached for . b)Defined as . c)Reached for . | |||||
d)Magnetic field corresponding to the vanishing conductance, . | |||||
e)Defined via the opening width ; i.e., . |
System | Defining parameters | System (sample) length, | No. of sites |
---|---|---|---|
Constriction with | nm | ||
zigzag edges | |||
Half-Corbino disk | nm | ||
Circular quantum dot | nm | ||
Circular quantum dot | nm | ||
with a circular hole | |||
—the distance between semi-infinite leads; —the distance between interfaces (given in parenthesis). | |||
b)Total no. of sites between the leads. (No. of sites with is given in parenthesis.) |
IV Results and discussion
In this Section, we confront the predictions from analytical theory for the charge-transfer cumulants (see Sec. II) with the results of computer simulations of electron transport in selected nanostructures in graphene shown in Fig. 1 (for the parameters, see Table 3). Since discrete structures carved out of a honeycomb lattice exhibit Fabry-Pérot type oscillations in all studied transport properties as a function of Fermi energy, and (typically) the higher the cumulant, the larger the oscillation magnitude, we limit the forthcoming discussion to the Landauer-Büttiker conductance () and the Fano factor (). It is also worth noting that the ratio of the former to the Sharvin conductance () accompanied by provides sufficient information to unambiguously identify one of the basic quantum transport regimes, see Table 2, if applicable.
IV.1 Tight-binding Hamiltonian
This part of the analysis starts from the tight-binding model of graphene, with Hamiltonian
(63) |
where the indices , run over sites in the honeycomb lattice of carbon atoms, and is the spin up/down orientation. The hopping-matrix elements are given by
(64) |
with eV. For the systems shown in Figs. 1(a), 1(d), and 1(e), the electrostatic potential energy varies only along the main axis. It equals , with eV, in the leads and raise to in the sample area. The abrupt potential increase at the sample-lead interface is smoothed over the length , according to the function
(65) |
The potential barrier, composed of two steps at and , namely
(66) |
is rectangular for [solid line in Fig. 1(b)], whereas it has a sinusoidal shape for [dashed line]. For the half-disk shown in Fig. 1(c), we simply take , where is the radius in polar coordinates, with the same function as in Eq. (66). The interface positions () coincide with the ends of the central (narrowest) part with parallel edges [see Fig. 1(a)], the inner/outer disk radii [Fig. 1(c)], or with the neckings limiting the dot region [Fig. 1(d,e)]. The remaining symbols in Eq. (63) are a creation (annihilation) operator for electron with spin at lattice site , () and the particle-number operator . (Since the Hamiltonian (63) can be represented as the sum of the two commuting terms, one for and the other for , it is sufficient to calculate the transport characteristics for one spin direction and to multiply the results by the degeneracy factor .)
In the following, we consider only for the constriction shown in Fig. 1(a); once the effect of the smooth potential barrier is identified, the discussion of the remaining systems concentrates on the case of (i.e., abrupt step).


IV.2 Constriction with zigzag edges
As a first example of the system, for which the analytical mode matching technique presented in Sec. II cannot be directly applied, we consider the constriction with zigzag edges, earlier considered as the valley Ryc07 or spin Wim08 ; Gru14 filter, depicted in Fig. 1(a). The central section of this system is an almost perfect square, with the length nm and the width nm (see also Table 3), attached to wedge-shaped electrodes that evolve into wide stripes with the width nm. Such a geometry is chosen to mimic the typical experimental situation, in which the nanostructure in graphene is contacted by much wider metallic leads Bao10 . Also, the potential step height, eV is not far from the results of some first-principles calculations for graphene-metal structures Gio08 ; Cus17 . Semi-infinite leads of a constant width play a role of the waveguides shown in Fig. 2; they can be divided into the repeating supercells in order to find the propagating modes numerically, by adapting the scheme developed by Ando for a square lattice And91 to the honeycomb lattice. For the potential profile given by Eqs. (65) and (66), the number of propagating modes (per one direction) is equal in the left and right leads, nmodfoo .
Since the central section of the system is bounded by two parallel interfaces separating weakly and heavily doped regions, one can expect that the key findings for a graphene strip in the sub-Sharvin regime, see Eqs. (44) and (51) still apply, at least for . However, the system width now varies with the position along the main axis, so the scattering cannot be described independently for each normal mode, as in Eq. (27). Instead, the mode mixing occurs, and — if scattering from the left is considered — we can define the transmission matrix , with and , and the reflection matrix , with , . The details of the calculations are presented in Appendix A; here we only mention that Eqs. (5) and (12) for measurable quantities remain valid, provided that the transmission probabilities are defined as eigenvalues of the matrix . Alternatively, one can express the Landauer-Büttiker conductance and the Fano factor in the basis-independent form, referring to the traces of the matrices and , namely
(67) | ||||
(68) |
The factor in Eq. (67) denotes the spin degeneracy. The valley degeneracy of the transmission eigenvalues is now only approximate, since the dispersion relation following from the Hamiltonian (63) is no longer perfectly conical, but shows the trigonal warping Kat20 . (For zigzag edges and electron doping, exact valley degeneracy occurs for all but one mode; for armchair edges the degeneracy is approximate for all modes Wak02 ; Wak10 .)
The results of our computer simulations are depicted by the thick colored lines in Fig. 6. They match the sub-Sharvin values (marked by black solid lines) for electron doping () and the abrupt potential step (). For hole doping () and the conductance is still close to as long as the number of propagating modes in the leads is sufficiently large [see the inset in Fig. 6(b)]. At the same time, the Fano factor is rather closer to the value of , which characterizes the symmetric cavity. In contrast, for smooth bariers () we have for and for (the conductance suppression due to the presence of two p-n junctions), as can be expected for the standard (i.e., Schödinger) ballistic system. At the same time, the Fano factor switches from (for ) to (for ). These findings are consistent with the results for smooth potential barriers and a strip with parallel edges, with mass confinement, presented in Ref. Ryc21b .
We see then that the constriction with zigzag edges carved out of a honeycomb lattice preserves all the key features of the idealized Dirac system studied previously.
IV.3 Half-disk and circular quantum dots
We now focus on the case of abrupt potential step () and consider the geometries for which the possible role of the edges is reduced (the half-Corbino disk) or amplified (circular quantum dot, without- or with a circular hole) compared to the constriction discussed above. The conductance and the Fano factor determined from Eqs. (67) and (68) after numerical calculation of the corresponding transmission matrix (see also Appendix A) are presented in Fig. 7.
In the half-disk case, see Figs. 7(a) and 7(d), the conductance (for ) remains in the interval (notice that the radii ratio is , and thus the relevant analytic approximations are given in Eq. (55) for the narrow-opening limit), with a tendency to approach the narrow-opening value with increasing . For , the conductance behavior is less clear, but the values of are still close to both and . In contrast, the Fano factor is close to the narrow-opening value of for both and , except in the small vicinity of the charge-neutrality point (), where it is noticeably closer to the pseudodiffusive value of .
For circular quantum dots Fabry-Pérot interference combined with scattering from irregular sample edges leads to much more pronounced oscillations of both and , discussed as functions of the Fermi energy, than in the case of a half-Corbino disk. In addition, the spectra presented in Figs. 7(b,c) for and 7(e,f) for suggest that the first charge-transfer characteristic (), discussed in isolation, may lead to the misidentification of the Sharvin or sub-Sharvin transport regime. Looking at the spectra, it is clear that the chaotic cavity (with ) is the closest of the simple models that captures key features of the circular quantum dot (both in the variant without- or with a hole), at least for higher electron or hole dopings. The conductance itself, related to the Sharvin value for , appears to be misleadingly close to in the absence of a hole, or to in the presence of a hole. (For , the suppression of due to p-n junctions occurs in both cases.)
Therefore, complex nanostructures with irregular edges may accidentally show some features of Sharvin (or sub-Sharvin) transport, but systematic discussion of quantum transport unveils the chaotic nature of the system.
V Conclusions
We have presented the analytical technique that allows one to calculate arbitrary charge transfer cumulant for doped graphene sample in two distinct physical situations: (i) two long and parallel abrupt interfaces separate the sample and the leads; (ii) a narrow circular interface governs transport through the much wider sample toward an external lead. In both cases, compact expressions are available for sufficiently high sample doping (infinite doping is assumed for the leads), for which multiple scattering between the interfaces can be taken into account, imposing the random phase each time the electron passes the sample area.
We have also reviewed the most common quantum transport regimes described in the literature, with their statistical distributions of transmission probabilities. Evidence for a novel sub-Sharvin transport regime in doped graphene is pointed out.
Next, the results of analytical considerations for idealized systems are compared with computer simulations of quantum transport for more realistic systems carved out of a honeycomb lattice. The effects of finite doping in the leads, smooth potential steps, trigonal warping, and irregular sample edges are included in our simulations. The results show that the main features of the analytical approach discussed in the first part, which defines the sub-Sharvin transport regime in graphene (with its variants for parallel interfaces and for the narrow-opening limit), are well reproduced in discrete systems on a honeycomb lattice, provided that the sample edges are straight and relatively short; i.e., with a total length comparable to or shorter than the total length of the sample-lead interfaces. In contrast, for systems with long and irregular edges, different charge-transfer cumulants may suggest different quantum transport regimes, making unambiguous classification difficult or impossible.
Although our paper focuses on graphene, we expect that the main effects will also occur in other two-dimensional crystals such as silicene, germanene, or stanene Eza15 ; Har19 . This prediction is based on the nature of the results presented, in particular the fact that the occurrence of the sub-Sharvin transport regime is related to the conical dispersion relation rather than to the transmission via evanescent waves, which is responsible for the graphene-specific phenomena that occur at the charge neutrality point.
Acknowledgements
The work was mainly completed during a sabbatical granted by the Jagiellonian University in the summer semester of 2023/24. We gratefully acknowledge Polish high-performance computing infrastructure PLGrid (HPC Center: ACK Cyfronet AGH) for providing computer facilities and support within computational grant no. PLG/2024/017208.

Appendix A Numerical mode-matching for the honeycomb lattice
A.1 The nanosystem
In the following, we show how the computational method originally developed by Ando for a square lattice And91 can be adapted to the honeycomb lattice. The key point is to notice that — when discussed on the level of tight-binding Hamiltonian (63) — a honeycomb lattice with nearest-neighbor hoppings only becomes equivalent to a square lattice with some bonds removed (see Fig. 8(a)). The reasoning presented here can, with simple modifications, be applied to a honeycomb, square, or even triangular lattice (the last case may be applicable when discussing the recently discovered two-dimensional form of gold, goldenene, see Ref. Kas24 ).
The nanosystem, shown schematically in Fig. 8(b), represents a generic case that is tractable within the Landauer-Büttiker formalism. Two semi-infinite leads are built from repeatable sections, allowing one to find the normal modes (propagating and evanescent) that define the basis for the scattering matrix — with moderate computational effort — via the secular equation (see next subsection); the central part may have an arbitrary shape, and thus the mode matching must be performed numerically on the lattice, which reduces to solving a sparse linear system of equations.
The details (and remaining assumptions) of both computational steps are given below.
A.2 Solutions in the leads
Keeping in mind the analogy between the honeycomb lattice and the square lattice mentioned above, we now consider an infinitely long wire with a width of sites, short sections of which are schematically shown in Figs. 8(b) (for zigzag edges) and 8(c) (for armchair edges). Quantum-mechanical equation of motion can be written as
(69) |
where
(70) |
is the -component wavefunction with the probability amplitudes corresponding to the sites in the and blocks; i.e, . It is worth noting that the lead width in physical units, , is related to as follows
(71) |
Therefore, can be interpreted as a circumference of a nanotube that can be created by connecting the edge sites at each section by additional (vertical) hopping.
For zigzag edges, and are matrices with the following block structure
(72) |
The block matrices contain values of the electostatic potential energy (i.e., for all sites) and vertical hopping elements depicted in Fig. 8; they are tridiagonal matrices with every second hopping removed, namely
(73) | ||||
(74) |
where , and is the Kronecker delta. The matrices and contain horizontal hoppings; for zero magnetic field, we have
(75) |
For a uniform magnetic field, the Peierls substitution can be applied; it reads (), with the flux per unit cell and the flux quantum. (The generalization for some geometric strains, leading to position-dependent hopping is also possible, provided that the invariance upon , see Eq. (69), is preserved.)
For armchair edges, the site-numbering scheme presented in Fig. 8(c) allows us to keep the block structure of Eq. (69), with
(76) |
Now, the matrices contain only diagonal elements,
(77) |
The remaining blocks, and , are given by
(78) | ||||
(79) |
where again.
Using the familiar Bloch ansatz,
(80) |
we arrive to the secular equation, for zigzag edges,
(81) |
where is the identity matrix and the eigenvalues are complex with (for propagating modes) or real with (for evanescent modes). For armchair edges, Eq. 81 holds with the following substitutions
(82) |
(If the generalization involving the dependence of the matrix elements on the position across the lead, such as the Peierls substitution mentioned above, is desired, one must double the number blocks in Eq. (76) by introducing ; however, such a case is beyond the scope of the present paper).
A.3 Basis for the scattering matrix
The secular equation derived above, given explicitly by Eq. (81) for zigzag edges (z.e.), with the necessary substitutions for armchair edges (a.e.) in Eq. (82), can be diagonalized numerically using standard software packages in order to find the full set of eigenvalues () with corresponding right-eigenvectors . We chose the double precision LAPACK routines dgemm and dgeev, see Ref. lapack99 .
To correctly define the scattering matrix, the propagating modes must be normalized so that they carry equal currents along the main axis of each lead (i.e., the -axis direction in Fig. 8).
In general, the total current incoming at the -th lattice site of the system described by the tight-binding Hamiltonian, see Eqs. (63) and (64) in the main text, is given by the time derivative of the charge,
(83) |
where runs over the nearest neighbors of , complex hopping (i.e., ) is allowed, and the spin is omitted for clarity. In turn, the current flowing from site to site is described by the quantum-mechanical operator
(84) |
Taking into account all the currents incoming and outgoing from a repeatable section () of the lead (see Fig. 8) via individual bonds, projected onto the -direction, bring us to the total -current operator , which — when acting on the right-eigenvector , see Eqs. (81) and (82), corresponding to the eigenvalue — is equivalent to
(85) |
or
(86) |
Subsequently, the normalized eigenvector can be written as
(87) |
and the sign
(88) |
indentifies the direction of propagation. For evanescent modes, the normalization is irrelevant, and we simply set
(89) |
In analogy to propagating modes, now identifies the evanescent mode decaying to the left, while identifies the evanescent mode decaying to the right. The real-space components of the normalized eigenvector for the -th section of the lead can be written as column vectors of the dimension , namely
(90) |
Now it is worth generalizing the discussion a bit by allowing that the left and right leads are not necessarily identical, i.e., may have different numbers of sites across, and , or different electostatic potential energy levels, and . In such a case, the secular equation, Eq. 81, appears in the two versions, with the block matrices replaced by
(91) |
leading to (a priori) different numbers of propagating modes per single direction, and . The number of left- (or right-) decaying evanescent modes is then given by
(92) |
Therefore, the number of elements (i.e., normalized eigenvectors) for the following four sets of (left-, right-) propagating and (left-, right-) decaying modes in a single electrode are as follows
(93) | ||||
(94) | ||||
For further considerations, we define the matrices and , in which the selected eigenvectors are stored column by column, i.e.,
(95) |
and
(96) |
In the above, the blocks and store the matrix rows and (respectively), for . These blocks correspond to the first sites (), and to the next sites () in the -th section of a single lead.
Now, the real-space components of an arbitrary wavefunction in the -th section of the (left,right) lead can be represented as
(97) | |||
(98) |
where each set of complex coefficients, , etc., corresponds to the matching set of eigenvectors in the matrix or . Parts of the matrices , , containing the columns corresponding to the coefficients , , read
(99) | ||||
(100) |
where each of the blocks , is a square matrix of dimension . The remaining columns, corresponding to the coefficients , define the matrices
(101) | ||||
(102) |
where the dimensions of the blocks , are , for .
A.4 The scattering problem
In the last part of this Appendix, we will use the notation relevant for the leads with zigzag edges (see Fig. 8). Nevertheless, if a version of the approach for the leads with armhair edges is desired, it can be easily generated by performing the substitutions given by Eq. 82. (We emphasize here that the central section of the system, described by the tight-binding Hamiltonian , may have an arbitrary shape, so that the irregular edges in the central section are also tractable within the presented approach.)
Using the definitions introduced in the previous subsections, we can now write down the quantum-mechanical equation of motion for the entire nanosystem as follows
(103) |
where we set for the terminal section of the left lead, and for the terminal section of the right lead. The unit matrix is omitted in expressions and ; the coefficients in each set, , etc., are now listed as columns; is a column vector that stores the wavefunction amplitudes for the central section.
In a case when the width of the central section, , exceeds the width of (at least) one of the leads, i.e., or (in addition, the leads may be vertically displaced, as in the nanosystem shown in Fig. 8), the blocks occurring near the upper-left or lower-right corner of the main matrix in Eq. (103) that connect the central section with the leads, namely, and with their hermitian conjugates , , are constructed in such a way that the original matrix is horizontally or vertically expanded and filled with zeros, namely
(104) |
where the elements in the original block match the existing bonds connecting the lattice sites. (For , we simply have and .)
In addition, Eq. (81) with substitutions given by Eq. (91) for the two leads, guarantees that
(105) | ||||
(106) |
allowing some further simplification of Eq. (103).
We now introduce the scattering matrix,
(107) |
defined in such a way that
(108) |
If we assume that the coefficients and are independent amplitudes of the incoming waves in the left and right leads, respectively, we can transform the linear system in Eq. (103) into a series of Kramers systems, one for each incoming mode ; namely,
(109) |
where is the set of wavefunctions for the central section, for , stored as columns. In the above, we used Eqs. (105), (106).
The rank of the main matrix in Eq. (109) is equal to , where denotes the number of vertical sections, each containing (up to) lattice sites. Although the shape of the central section is — in principle — irregular, a site-numbering scheme analogous to that used for the leads can be applied. In case of a considerable number of disconnected lattice sites, an additional enumeration for the central part can be applied to eliminate the matrix elements corresponding to disconnected sites and to reduce the rank of the main matrix.
For the total number of sites , such as in the numerical examples presented in the main text, one can use standard linear-algebra software to solve the linear system given by Eq. (109) numerically and find the unknowns, including all elements of the scattering matrix (107). We used the double precision LAPACK routine zgbsv, see Ref. lapack99 , which employs the band storage scheme for the main matrix with subdiagonals and superdiagonals. Assuming that the proportions of the middle section are not very different from those of the square, i.e., (with the total number of sites ) we can estimate the computational complexity to be , which is equivalent to the complexity of the more commonly used recursive Green’s function method Lew13 .
Since the normal modes in the left and right leads are calculated independently, we can validate the final numerical output by checking the unitarity condition for the scattering matrix, . The deviation from unitarity is quantified by
(110) |
In the numerical problems considered here, the above does not exceed ; the maximal value is approached for the half-disk system, for which the number of propagating modes in the right (i.e., the wider) lead is for the energy range of .
If more than two leads contact the central system, a generalization is straightforward, provided we have a group of (one or more) parallel left leads and a group of parallel right leads. In such a case, the block structure of Eq. (109) is preserved, and it is only necessary to modify the contents of the blocks connecting the central section and the leads accordingly. For some more complex cases, relevant graph algorithms are implemented in the KWANT package Gro14 .
For much larger systems, the direct lattice approach presented above must be replaced by the method truncating the wavefunction within orthogonal polynomials, such as recently implemented in the KITE software Joa20 .
References
- (1) M. J. Duff, L. B. Okun, and G. Veneziano, Trialogue on the number of fundamental constants, J. High Energy Phys. 03, 023 (2002).
- (2) J.-M. Lévy-Leblond, On the Conceptual Nature of the Physical Constants, in The Reform of the International System of Units (SI): Philosophical, Historical and Sociological Issues (Routledge, London, UK, 2019), pp. 125-149. DOI: https://doi.org/10.4324/9781351048989.
- (3) T. Ando, Y. Matsumoto, and Y. Uemura, Theory of Hall Effect in a Two-Dimensional Electron System, J. Phys. Soc. Jpn. 39, 279 (1975).
- (4) K. v. Klitzing, G. Dorda, and M. Pepper, New Method for High-Accuracy Determination of the Fine-Structure Constant Based on Quantized Hall Resistance, Phys. Rev. Lett. 45, 494 (1980).
- (5) R. B. Laughlin, Quantized Hall conductivity in two dimensions, Phys. Rev. B 23, 5632(R) (1981).
- (6) D. C. Tsui, H. L. Stormer, and A. C. Gossard, Two-Dimensional Magnetotransport in the Extreme Quantum Limit, Phys. Rev. Lett. 48, 1559 (1982).
- (7) B. D. Josephson, The discovery of tunneling supercurrents, Rev. Mod. Phys. 46, 251 (1974).
- (8) S. Payagala, A. Rigosi, A. Panna, A. Pollarolo, M. Kruskopf, S. Schlamminger, D. Jarrett, R. Brown, R. Elmquist, D. Brown, and D. Newell, Comparison between Graphene and GaAs Quantized Hall Devices with a Dual Probe, IEEE Trans. Instrum. Meas. 69, 9374 (2020).
- (9) Y. Tang, J. Wachter, A. Rufenacht, G. FitzPatrick, and S. Benz, 10 V Programmable Josephson Voltage Standard and its Application in Direct Comparison with the Conventional Josephson Voltage Standard, IEEE Trans. Instrum. Meas. 64, 3458 (2015).
- (10) K. Novoselov, A. Geim, S. Morozov, D. Jiang, M. I. Katsnelson, I. V. Grigorieva, S. V. Dubonos, and A. A. Firsov Two-dimensional gas of massless Dirac fermions in graphene, Nature 438, 197 (2005).
- (11) Y. Zhang, Y.-W. Tan, H. L. Stormer, and P. Kim, Experimental observation of the quantum Hall effect and Berry’s phase in graphene, Nature 438, 201 (2005).
- (12) M. Titov and C. W. J. Beenakker, Josephson effect in ballistic graphene, Phys. Rev. B 74, 041401(R) (2006).
- (13) S. Salim, R. Marathe, and S. Gosh, Revisiting Andreev processes in superconductor-graphene-superconductor (SGS) Josephson junctions: comparison with experimental results, Phys. Scr. 98, 065935 (2023).
- (14) M. I. Katsnelson, Zitterbewegung, chirality, and minimal conductivity in graphene, Eur. Phys. J. B 51, 157 (2006).
- (15) M. I. Katsnelson, The Physics of Graphene, 2nd ed. (Cambridge University Press, Cambridge, UK, 2020); Chapter 3.
- (16) R. R. Nair, P. Blake, A. N. Grigorenko, K. S. Novoselov, T. J. Booth, T. Stauber, N. M. R. Peres, and A. K. Geim, Fine Structure Constant Defines Visual Transparency of Graphene, Science 320, 1308 (2008).
- (17) H. S. Skulason, P. E. Gaskell, and T. Szkopek, Optical reflection and transmission properties of exfoliated graphite from a graphene monolayer to several hundred graphene layers, Nanotechnology 21, 295709 (2010).
- (18) J. Tworzydło, B. Trauzettel, M. Titov, A. Rycerz, and C. W. J. Beenakker, Sub-Poissonian shot noise in graphene, Phys. Rev. Lett. 96, 246802 (2006).
- (19) R. Danneau, F. Wu, M. F. Craciun, S. Russo, M. Y. Tomi, J. Salmilehto, A. F. Morpurgo, and P. J. Hakonen, Shot Noise in Ballistic Graphene, Phys. Rev. Lett. 100, 196802 (2008).
- (20) A. Rycerz, P. Recher, and M. Wimmer, Conformal mapping and shot noise in graphene, Phys. Rev. B 80, 125417 (2009).
- (21) A. Laitinen, G. S. Paraoanu, M. Oksanen, M. F. Craciun, S. Russo, E. Sonin, and P. Hakonen, Contact doping, Klein tunneling, and asymmetry of shot noise in suspended graphene, Phys. Rev. B 93, 115413 (2016).
- (22) H. Yoshino and K. Murata, Significant Enhancement of Electronic Thermal Conductivity of Two-Dimensional Zero-Gap Systems by Bipolar-Diffusion Effect. J. Phys. Soc. Jpn. 84, 024601 (2015).
- (23) J. Crossno, J. K. Shi, K. Wang, X. Liu, A. Harzheim, A. Lucas, S. Sachdev, P. Kim, T. Taniguchi, K. Watanabe, T. A. Ohki, and K. C. Fong, Observation of the Dirac fluid and the breakdown of the Wiedemann-Franz law in graphene, Science 351, 1058 (2016).
- (24) A. Rycerz, Wiedemann–Franz Law for Massless Dirac Fermions with Implications for Graphene, Materials 14, 2704 (2021).
- (25) Y.-T. Tu and S. Das Sarma, Wiedemann-Franz law in graphene, Phys. Rev. B 107, 085401 (2023).
- (26) G. W. Semenoff, Condensed-Matter Simulation of a Three-Dimensional Anomaly, Phys. Rev. Lett. 53, 2449 (1984).
- (27) D. P. Di Vincenzo, E. J. Mele, Self-consistent effective-mass theory for intralayer screening in graphite intercalation compounds, Phys. Rev. B 29, 1685 (1984).
- (28) A. Rycerz and P. Witkowski, Sub-Sharvin conductance and enhanced shot noise in doped graphene, Phys. Rev. B 104, 165413 (2021).
- (29) A. Rycerz and P. Witkowski, Theory of sub-Sharvin charge transport in graphene disks, Phys. Rev. B 106, 155428 (2022).
- (30) Yu. V. Sharvin, A possible method for studying Fermi surfaces, Zh. Eksp. Teor. Fiz. 48, 984 (1965) [Sov. Phys. JETP 21, 655 (1965)].
- (31) B. J. van Wees, H. van Houten, C. W. J. Beenakker, J. G. Williamson, L. P. Kouwenhoven, D. van der Marel, and C. T. Foxon, Quantized conductance of point contacts in a two-dimensional electron gas, Phys. Rev. Lett. 60, 848 (1988).
- (32) L. I. Glazman, G. B. Lesovik, D. E. Khmel’nitskii, and R. I. Shekhter, Reflectionless quantum transport and fundamental ballistic-resistance steps in microscopic constrictions, Pi’sma Zh. Eksp. Teor. Fiz. 48, 218 (1988) [JETP Lett. 48, 238 (1988).]
- (33) For the disk geometry, the width needs to be replaced with the inner lead circumfence .
- (34) A. Rycerz, K. Rycerz, and P. Witkowski, Thermoelectric Properties of the Corbino Disk in Graphene, Materials 16, 4250 (2023).
- (35) A. Rycerz, K. Rycerz, and P. Witkowski, Sub-Sharvin Conductance and Incoherent Shot-Noise in Graphene Disks at Magnetic Field, Materials 17, 3067 (2024).
- (36) S. Das Sarma and K. Yang, The enigma of the nu=0 quantum Hall effect in graphene, Solid State Commun. 149, 1502 (2009).
- (37) A. Rycerz, Magnetoconductance of the Corbino disk in graphene, Phys. Rev. B 81, 121404(R) (2010).
- (38) Y. Zeng, J. I. A. Li, S. A. Dietrich, O. M. Ghosh, K. Watanabe, T. Taniguchi, J. Hone, and C. R. Dean, High-Quality Magnetotransport in Graphene Using the Edge-Free Corbino Geometry, Phys. Rev. Lett. 122, 137701 (2019).
- (39) D. Suszalski, G. Rut, and A. Rycerz, Mesoscopic valley filter in graphene Corbino disk containing a p-n junction J. Phys. Mater. 3, 015006 (2020).
- (40) M. Kamada, V. Gall, J. Sarkar, M. Kumar, A. Laitinen, I. Gornyi, and P. Hakonen, Strong magnetoresistance in a graphene Corbino disk at low magnetic fields, Phys. Rev. B 104, 115432 (2021).
- (41) C. Kittel, Introduction to Solid State Physics, 8th ed. (John Willey and Sons: New York, NY, USA, 2005); Chapter 6.
- (42) C. T. White and T. N. Todorov, Carbon nanotubes as long ballistic conductors, Nature 393, 240 (1998).
- (43) N. Agraït, A. Levy Yeyati, J. M. van Ruitenbeek, Quantum properties of atomic-sized conductors, Phys. Rep. 377, 81 (2003).
- (44) R. Landauer, Spatial Variation of Currents and Fields Due to Localized Scatterers in Metallic Conduction, IBM J. Res. Dev. 1, 223 (1957).
- (45) M. Büttiker, Y. Imry, R. Landauer, and S. Pinhas, Generalized many-channel conductance formula with application to small rings, Phys. Rev. B 31, 6207 (1985).
- (46) C. Kumar, J. Birkbeck, J. A. Sulpizio, D. J. Perello, T. Taniguchi, K. Watanabe, O. Reuven, T. Scaffidi, A. Stern, A. K. Geim, and S. Ilani, Imaging hydrodynamic electrons flowing without Landauer-Sharvin resistance, Nature 609, 276 (2022).
- (47) Y. Cao, V. Fatemi, A. Demir, S. Fang, S. L. Tomarken, J. Y. Luo, J. D. Sanchez-Yamagishi, K. Watanabe, T. Taniguchi, E. Kaxiras, R. C. Ashoori, and P. Jarillo-Herrero, Correlated insulator behaviour at half-filling in magic-angle graphene superlattices, Nature 556, 80 (2018).
- (48) Y. Cao, V. Fatemi, S. Fang, K. Watanabe, T. Taniguchi, E. Kaxiras, and P. Jarillo-Herrero, Unconventional superconductivity in magic-angle graphene superlattices, Nature 556, 43 (2018).
- (49) M. Fidrysiak, M. Zegrodnik, and J. Spałek, Unconventional topological superconductivity and phase diagram for an effective two-orbital model as applied to twisted bilayer graphene, Phys. Rev. B 98, 085436 (2018).
- (50) Y. V. Nazarov and Y. M. Blanter, Quantum Transport: Introduction to Nanoscience (Cambridge University Press, Cambridge, UK, 2009); Chapter 1.
- (51) Strictly speaking, these solutions take the form of linear combinations of two plane waves (with positive and negative ), producing standing waves in the direction, but this detail is irrelevant to the calculations presented below.
- (52) E. C. Kemble, A Contribution to the Theory of the B. W. K. Method, Phys. Rev. 48, 549 (1935).
- (53) N. Takimoto, S. Yamamura, and K. Funayama, Generation of Non-Equilibrium 1/f Temperature Noise, J. Phys. Soc. Jpn. 62, 3077 (1993).
- (54) K. Schönhammer, Full counting statistics for noninteracting fermions: Exact results and the Levitov-Lesovik formula, Phys. Rev. B 75, 205329 (2007).
- (55) It should be noted that Eq. (20) is used for the -valley neighborhood in the dispersion relation; for the -valley the matrix should be replaced by .
- (56) M. V. Berry and R. J. Mondragon, Neutrino billiards: time-reversal symmetry-breaking without magnetic fields, Proc. R. Soc. Lond. A 412, 53 (1987).
- (57) T. R. Robinson, On Klein tunneling in graphene, Am. J. Phys. 80, 141 (2012).
- (58) C. Gutiérrez, L. Brown, C. J. Kim, J. Park, and A. N. Pasupathy, Klein tunnelling and electron trapping in nanometre-scale graphene quantum dots, Nat. Phys. 12, 1069 (2016).
- (59) Similar to the Schrödinger equation case, the wave function matching conditions are a consequence of charge conservation. To obtain the current density operator, in the Hamiltonian operator on the left side of Eq. (20), , we substitute , and then differentiate the resulting operator with respect to the vector potential . In effect, we obtain , and therefore the continuity of the current corresponds to the continuity of the components of the wave function (operator does not involve differentiation over coordinates).
- (60) S. Datta, Electronic Transport in Mesoscopic Systems (Cambridge University Press, Cambridge, UK 1997); Chapter 3.
- (61) A. Schuessler, P. M. Ostrovsky, I. V. Gornyi, and A. D. Mirlin, Full counting statistics in disordered graphene at Dirac point: From ballistics to diffusion, Phys. Rev. B 82, 085419 (2010).
- (62) I. S. Gradshteyn and I. M. Ryzhik, Table of Integrals, Series, and Products, Seventh Edition (Academic Press, New York, 2007); Eq. 2.553.3.
- (63) F. Miao, S. Wijeratne, Y. Zhang, U. C. Coscun, W. Bao, C. N. Lau, Phase Coherent Transport in Graphene Quantum Billiards, Science 317, 1530 (2007).
- (64) M. Abramowitz and I. A. Stegun, eds., Handbook of Mathematical Functions (Dover Publications, Inc., New York, 1965); Chapter 15.
- (65) L. DiCarlo, J. R. Williams, Y. Zhang, D. T. McClure, C. M. Marcus, Shot Noise in Graphene, Phys. Rev. Lett. 100, 156801 (2008).
- (66) M. I. Katsnelson and F. Guinea, Transport through evanescent waves in ballistic graphene quantum dots, Phys. Rev. B 78, 075417 (2008).
- (67) A. Rycerz and D. Suszalski, Graphene disk in a solenoid magnetic potential: Aharonov-Bohm effect without a two-slit-like setup, Phys. Rev. B 101, 245429 (2020).
- (68) C. W. J. Beenakker and M. Büttiker, Suppression of shot noise in metallic diffusive conductors, Phys. Rev. B 46, 1889(R) (1992).
- (69) K. E. Nagaev, On the shot noise in dirty metal contacts, Phys. Lett. A 169, 103 (1992).
- (70) C. W. J. Beenakker, Random-matrix theory of quantum transport, Rev. Mod. Phys. 69, 731 (1997).
- (71) A. Rycerz, J. Tworzydło, and C. W. J. Beenakker, Valley filter and valley valve in graphene, Nat. Phys. 3, 172 (2007).
- (72) M. Wimmer, İ. Adagideli, S. Berber, D. Tománek, and K. Richter, Spin Currents in Rough Graphene Nanoribbons: Universal Fluctuations and Spin Injection, Phys. Rev. Lett. 100, 177207 (2008).
- (73) M. M. Grujić, M. Ž. Tadić, and F. M. Peeters, Spin-Valley Filtering in Strained Graphene Structures with Artificially Induced Carrier Mass and Spin-Orbit Coupling, Phys. Rev. Lett. 113, 046601 (2014).
- (74) W. Bao, G. Liu, Z. Zhao, H. Zhang, D. Yan, A. Deshpande, B. J. LeRoy, and C. N. Lau, Lithography-free Fabrication of High Quality Substrate-supported and Freestanding Graphene devices, Nano Res. 3, 98 (2010).
- (75) G. Giovannetti, P. A. Khomyakov, G. Brocks, V. M. Karpan, J. van den Brink, and P. J. Kelly, Doping Graphene with Metal Contacts, Phys. Rev. Lett. 101, 026803 (2008).
- (76) T. Cusati, G. Fiori, A. Gahoi, V. Passi, M. C. Lemme, A. Fortunelli, and G. Iannaccone, Electrical properties of graphene-metal contacts, Sci. Rep. 7, 5109 (2017).
- (77) T. Ando, Quantum point contacts in magnetic fields, Phys. Rev. B 44, 8017 (1991).
- (78) It can be further approximated as with , giving for and ; the actual number of propagating modes (if ) due to the trigonal warping.
- (79) K. Wakabayashi and T. Aoki, Electrical conductance of zigzag nanographite ribbons with locally applied gate voltage, Int. J. Mod. Phys. B 16, 4897 (2002).
- (80) K. Wakabayashi, K. Sasaki, T. Nakanishi, and T. Enoki, Electronic states of graphene nanoribbons and analytical solutions, Sci. Technol. Adv. Mater. 11 054504 (2010).
- (81) M. Ezawa, Monolayer Topological Insulators: Silicene, Germanene, and Stanene, J. Phys. Soc. Jpn. 84, 121003 (2015).
- (82) T. Hartman and Z. Sofer, Beyond Graphene: Chemistry of Group 14 Graphene Analogues: Silicene, Germanene, and Stanene, ACS Nano 13, 8566 (2019).
- (83) S. Kashiwaya, Y. Shi, J. Lu, D. G. Sangiovanni, G. Greczynski, M. Magnuson, M. Andersson, J. Rosen, and L. Hultman, Synthesis of goldene comprising single-atom layer gold, Nat. Synth. 3, 744 (2024).
- (84) E. Anderson, Z. Bai, C. Bischof, S. Blackford, J. Demmel, J. Dongarra, J. D. Croz, A. Greenbaum, S. Hammarling, A. McKenney, et al., LAPACK Users’ Guide, Third Edition (Society for Industrial and Applied Mathematics, Philadelphia, PA, USA, 1999).
- (85) C. H. Lewenkopf, E. R. Mucciolo, The recursive Green’s function method for graphene, J. Comput. Eletron. 12, 203 (2013).
- (86) C. W. Groth, M. Wimmer, A. R. Akhmerov, and X. Waintal, Kwant: a software package for quantum transport, New J. Phys. 16, 063065 (2014).
- (87) S. M. João, M. Anđelković, L. Covaci, T. G. Rappoport, J. M. V. P. Lopes, and A. Ferreira, KITE: high-performance accurate modelling of electronic structure and response functions of large molecules, disordered crystals and heterostructures, R. Soc. Open Sci. 7, 191809 (2020).