Skip to main content
Log in

Quantum Transmission Conditions for Diffusive Transport in Graphene with Steep Potentials

  • Published:
Journal of Statistical Physics Aims and scope Submit manuscript

Abstract

We present a formal derivation of a drift-diffusion model for stationary electron transport in graphene, in presence of sharp potential profiles, such as barriers and steps. Assuming the electric potential to have steep variations within a strip of vanishing width on a macroscopic scale, such strip is viewed as a quantum interface that couples the classical regions at its left and right sides. In the two classical regions, where the potential is assumed to be smooth, electron and hole transport is described in terms of semiclassical kinetic equations. The diffusive limit of the kinetic model is derived by means of a Hilbert expansion and a boundary layer analysis, and consists of drift-diffusion equations in the classical regions, coupled by quantum diffusive transmission conditions through the interface. The boundary layer analysis leads to the discussion of a four-fold Milne (half-space, half-range) transport problem.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3

Similar content being viewed by others

Notes

  1. According to the terminology adopted, e.g., in [1], we call “semiclassical” a classical transport (or Boltzmann) equation where elements of quantum nature are retained, e.g. a non-parabolic dispersion relation.

  2. We remark that we are using the potential energy V instead of the electric potential \(-V/q\) (where \(q>0\) is the elementary charge).

  3. Note that here we are using non-dimensional chemical potentials, while the dimensional chemical potentials, that have the dimensions of a energy, are given by \(\beta ^{-1}A_s\).

  4. The normalization constant is required in order to get the the correct moments of a non-dimensional Wigner function [3].

  5. Note that the constant \(C_i\) is finite because conservation of energy holds with different signs of s and \(s'\) only in a finite energy interval, which corresponds to a bounded region in \({\varvec{p}}\)-space.

  6. Since we are using dimensionless phase-space distributions, the physical dimensions of \(n_s^{i,\infty }\) are actually those of a frequency.

  7. Except positivity, that is not guaranteed (and not required) here.

References

  1. Ashcroft, N.W., Mermin, N.D.: Solid State Physics. Saunders College Publishing, Philadelphia (1976)

    MATH  Google Scholar 

  2. Bardos, C., Santos, R., Sentis, R.: Diffusion approximation and the computation of the critical size. T. Am. Math. Soc. 284, 617–649 (1984)

    Article  MathSciNet  MATH  Google Scholar 

  3. Barletti, L.: Hydrodynamic equations for electrons in graphene obtained from the maximum entropy principle. J. Math. Phys. 55, 083303 (2014)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  4. Barletti, L.: Hydrodynamic equations for an electron gas in graphene. J. Math. Ind. 6, 7 (2016)

    Article  MathSciNet  Google Scholar 

  5. Barletti, L., Negulescu, C.: Hybrid classical-quantum models for charge transport in graphene with sharp potentials. J. Comput. Theor. Transp. 46, 159–175 (2017)

    Article  MathSciNet  Google Scholar 

  6. Barletti, L., Frosali, G., Morandi, O.: Kinetic and hydrodynamic models for multi-band quantum transport in crystals. In: Ehrhardt, M., Koprucki, T. (eds.) Multi-band Effective Mass Approximations: Advanced Mathematical Models and Numerical Techniques, pp. 3–56. Springer, Heidelberg (2014)

    Google Scholar 

  7. Ben Abdallah, N.: A hybrid kinetic-quantum model for stationary electron transport. J. Stat. Phys. 90, 627–662 (1998)

    Article  MathSciNet  MATH  Google Scholar 

  8. Ben, Abdallah N., Degond, P., Gamba, I.: Coupling one-dimensional time-dependent classical and quantum transport models. J. Math. Phys. 43, 1–24 (2002)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  9. Borysenko, K.M., Mullen, J.T., Barry, E.A., Paul, S., Semenov, Y.G., Zavada, J.M., Buongiorno Nardelli, M., Kim, K.W.: First-principles analysis of electron-phonon interactions in graphene. Phys. Rev. B 81, 121412(R) (2010)

    Article  ADS  Google Scholar 

  10. Castro Neto, A.H., Guinea, F., Peres, N.M.R., Novoselov, K.S., Geim, A.K.: The electronic properties of graphene. Rev. Mod. Phys. 81, 109–162 (2009)

    Article  ADS  Google Scholar 

  11. Cheianov, V.V., Fal’ko, V., Altshuler, B.L.: The focusing of electron flow and a Veselago lens in graphene. Science 315, 1252–1255 (2007)

    Article  ADS  Google Scholar 

  12. Degond, P.: Macroscopic limits of the Boltzmann equation: a review. In: Degond, P., Pareschi, L., Russo, G. (eds.) Modeling and Computational Methods for Kinetic Equations, pp. 3–57. Birkhäuser, Basel (2004)

    Chapter  Google Scholar 

  13. Degond, P., El Ayyadi, A.: A coupled Schrödinger drift-diffusion model for quantum semiconductor device simulations. J. Comput. Phys. 181, 222–259 (2002)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  14. Degond, P., Schmeiser, C.: Macroscopic models for semiconductor heterostructures. J. Math. Phys. 39, 4634–4663 (1998)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  15. Duderstadt, J.J., Martin, W.R.: Transport Theory. Wiley, New York (1979)

    MATH  Google Scholar 

  16. Golse, F., Klar, A.: A numerical method for computing asymptotic states and outgoing distributions for kinetic linear half-space problems. J. Stat. Phys. 80, 1033–1061 (1995)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  17. Katsnelson, M.I., Novoselov, K.S., Geim, A.K.: Chiral tunnelling and the Klein paradox in graphene. Nat. Phys. 2, 620–625 (2006)

    Article  Google Scholar 

  18. Lee, G.H., Park, G.H., Lee, H.J.: Observation of negative refraction of Dirac fermions in graphene. Nat. Phys. 11, 925–929 (2015)

    Article  Google Scholar 

  19. Lejarreta, J.D., Fuentevilla, C.H., Diez, E., Cerveró, J.M.: An exact transmission coefficient with one and two barriers in graphene. J. Phys. A 46, 155304 (2013)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  20. Majorana, A., Mascali, G., Romano, V.: Charge transport and mobility in monolayer graphene. J. Math. Ind. 7, 4 (2017)

    Article  MathSciNet  Google Scholar 

  21. Morandi, O.: Wigner-function formalism applied to the Zener band transition in a semiconductor. Phys. Rev. B 80, 024301 (2009)

    Article  ADS  Google Scholar 

  22. Poupaud, F.: Diffusion approximation of the linear semiconductor Boltzmann equation: analysis of boundary layers. Asymptot. Anal. 4, 293–317 (1991)

    MathSciNet  MATH  Google Scholar 

  23. Slonczewski, J.C., Weiss, P.R.: Band structure of graphite. Phys. Rev. 109, 272–279 (1958)

    Article  ADS  Google Scholar 

  24. Young, A.F., Kim, P.: Quantum interference and Klein tunnelling in graphene heterojunctions. Nat. Phys. 5, 222–226 (2009)

    Article  Google Scholar 

Download references

Acknowledgements

Support is acknowledged from the Italian-French project Projet International de Coopration Scientifique (PICS) “MANUS—Modelling and Numerics for Spintronics and Graphene” (Ref. PICS07373).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Luigi Barletti.

Appendix: Proof of Theorem 4.1

Appendix: Proof of Theorem 4.1

In order to streamline the notation, let us put

$$\begin{aligned} L^i(\varvec{z}) = L^i_s({\varvec{p}}) := F'_{n^i_s}({\varvec{p}}), \end{aligned}$$
(A.1)

and rewrite the Milne problem (4.29) accordingly:

$$\begin{aligned} \left\{ \begin{aligned}&\mu \frac{\partial \theta ^i}{\partial \xi } = L^i {\big \langle {\theta ^i} \big \rangle } - \theta ^i,&(-1)^i&\xi >0, \\&\theta ^i_\mathrm {in}- \mathcal {K}^i\big (\theta ^i_\mathrm {out},\theta ^j_\mathrm {out}\big ) = G^i_\mathrm {in}- \mathcal {K}^i\big (G^i_\mathrm {out},G^j_\mathrm {out}\big ),&\xi =0. \end{aligned} \right. \end{aligned}$$
(A.2)

We recall that in problem (A.2) the y-variable is just a parameter, which shall be omitted throughout the proof.

The proof is inspired by the ideas of Ref. [14] and is divided into three steps for the reader’s convenience.

\({ {{Step 1: reduction~to~a }~{\vert {{\varvec{p}}} \vert }{-averaged~problem.}}}\) Let us consider the uncoupled version of (A.2), with assigned inflows \(g^i\):

$$\begin{aligned} \left\{ \begin{aligned}&\mu \frac{\partial \theta ^i}{\partial \xi }(\xi ,\varvec{z}) = L^i (\varvec{z}){\big \langle {\theta ^i} \big \rangle } - \theta ^i(\xi ,\varvec{z}),&\quad (-1)^i&\xi >0,\quad \varvec{z}\in \varTheta ,\\&\theta ^i_\mathrm {in}(\varvec{z}) = g^i(\varvec{z}),&\xi =0, \quad \varvec{z}\in \varTheta ^i_\mathrm {in}\end{aligned} \right. \end{aligned}$$
(A.3)

(recall definitions (A.1), (3.23) and (4.19)). We introduce the \({\vert {{\varvec{p}}} \vert }\)-average

$$\begin{aligned} {\tilde{\theta }}^i(\xi ,\varphi ,s) := \frac{1}{2\pi \hbar ^2} \int _0^{+\infty } \theta ^i(\xi ,{\vert {{\varvec{p}}} \vert }\cos \varphi ,{\vert {{\varvec{p}}} \vert }\sin \varphi ,s)\,{\vert {{\varvec{p}}} \vert }\,d{\vert {{\varvec{p}}} \vert }, \end{aligned}$$
(A.4)

where the normalisation constant is chosen so that

$$\begin{aligned} {\big \langle {\theta ^i} \big \rangle }(\xi ,s) = \frac{1}{2\pi }\int _0^{2\pi } {\tilde{\theta }}^i(\xi ,\varphi ,s) \,d\varphi \quad \text {and} \quad {\tilde{L}}^i_s = 1, \end{aligned}$$

where \({\left\langle \theta ^i \right\rangle }\) is defined in (3.12). We introduce the analogous of the sets (3.22) for the averaged quantities:

$$\begin{aligned} {\tilde{\varTheta }} := [0,2\pi ) \times \{-1,+1\}, \quad {\tilde{\varTheta }}^i_{\mathrm {in}/\mathrm {out}} := \Big \{ (\varphi ,s) \in \varTheta \mid (-1)^i \cos \varphi \gtrless 0 \Big \} \end{aligned}$$
(A.5)

and extend, in the obvious way, to \({\tilde{\theta }}^i\) the notations \({\tilde{\theta }}^i_\mathrm {in}\) and \({\tilde{\theta }}^i_\mathrm {out}\). Taking the \({\vert {{\varvec{p}}} \vert }\)-average of (A.3), we obtain that \({\tilde{\theta }}^i_s\) satisfies

$$\begin{aligned} \left\{ \begin{aligned}&\mu \frac{\partial {\tilde{\theta }}^i}{\partial \xi }(\xi ,\varphi ,s) = \frac{1}{2\pi } \int _0^{2\pi } {\tilde{\theta }}^i(\xi ,\varphi ,s)\,d\varphi - {\tilde{\theta }}^i(\xi ,\varphi ,s),&\ (-1)^i&\xi >0,\ (\varphi ,s) \in {\tilde{\varTheta }},\\&{\tilde{\theta }}^i_\mathrm {in}(\varphi ,s) = {\tilde{g}}^i(\varphi ,s),&\xi =0,\ (\varphi ,s) \in {\tilde{\varTheta }}^i_\mathrm {in}, \end{aligned} \right. \end{aligned}$$
(A.6)

where we recall that

$$\begin{aligned} \mu = \frac{cp_x}{{\vert {{\varvec{p}}} \vert }} = c\cos \varphi . \end{aligned}$$

Conversely, it is easy to check that, if \({\tilde{\theta }}^i\) is a solution of (A.6), then

$$\begin{aligned} \theta ^i := L^i{\tilde{\theta }}^i + \left\{ \begin{aligned}&\mathrm {e}^{-\xi /\mu }\left( g^i - L^i {\tilde{g}}^i \right) ,&(-1)^i \mu >0,\\&0,&(-1)^i \mu <0, \end{aligned} \right. \end{aligned}$$
(A.7)

is solution of (A.3). The equations (A.6) are four (\(s = \pm 1\), \(i = 1,2\)) independent Milne problems having the form of a Milne problem for neutron transport, to the extent that the kernel of the collision operator coincides with the functions that are constant with respect to \(\varphi \) [2, 15]. About such problem the following facts are known [2, 14, 22]:

  1. (i)

    If \({\tilde{g}}^i \in \mathrm{L}^\infty ({\tilde{\varTheta }}^i_\mathrm {in})\), the solution \({\tilde{\theta }}^i\) to problem (A.6) exists and is unique in \(\mathrm{L}^\infty \big ( (-1)^i[0,+\infty )\times {\tilde{\varTheta }} \big )\). Moreover, one has the positivity, i.e. \({\tilde{\theta }}^i \ge 0\) if \({\tilde{g}}^i \ge 0\).

  2. (ii)

    A constant \(n_s^{i,\infty }\) (depending on \({\tilde{g}}^i\)) exists such that \({\tilde{\theta }}^i(\xi ,\varphi ,s) \rightarrow n_s^{i,\infty }\), as \(\xi \rightarrow (-1)^i\infty \), and the convergence is exponentially fast; in particular

    $$\begin{aligned} \Big \vert \int _0^{2\pi } {\tilde{\theta }}^i(\xi ,\varphi ,s)\, d\varphi - n_s^{i,\infty } \Big \vert \le C\mathrm {e}^{-\alpha {\vert {\xi } \vert }}, \end{aligned}$$

    for some constants \(C>0\) and \(\alpha >0\). Moreover, \(n^{i,\infty } \ge 0\) if \({\tilde{g}}^i \ge 0\).

  3. (iii)

    The Albedo operator, that associates the inflow to the outflow, i.e. \( {\tilde{\theta }}^i_\mathrm {in}\equiv {\tilde{g}}^i \mapsto {\tilde{\theta }}^i_\mathrm {out}\), is a compact linear operator from \(\mathrm{L}^\infty ({\tilde{\varTheta }}^i_\mathrm {in})\) to \(\mathrm{L}^\infty ({\tilde{\varTheta }}^i_\mathrm {out})\).

Step 2: formulation as a Fredholm problem. Thanks to the explicit formula (A.7), it is easy to extend the above results to the \({\vert {{\varvec{p}}} \vert }\)-dependent problem (A.3). Let us define the weighted spaces

$$\begin{aligned} X^i := \mathrm{L}^\infty \Big ( \varTheta , (L^i)^{-1} d{\varvec{p}}\Big ), \ X^i_\mathrm {in}:= \mathrm{L}^\infty \Big (\varTheta ^i_\mathrm {in}, (L^i)^{-1} d{\varvec{p}}\Big ), \ X^i_\mathrm {out}:= \mathrm{L}^\infty \Big (\varTheta ^i_\mathrm {out}, (L^i)^{-1} d{\varvec{p}}\Big ). \end{aligned}$$

Note that \(g \in X^i\) implies that \(g \in \mathrm{L}^p(\varTheta )\) for all \(p \in [0,\infty ]\). Then, from the results of Step 1, we have the following facts about problem (A.3):

  1. (i)

    If \(g^i \in X^i_\mathrm {in}\), the solution \(\theta ^i\) to problem (A.3) exists and is unique in the space \(\mathrm{L}^\infty \left( (-1)^i[0,+\infty ), X^i \right) \). Moreover, \(\theta ^i \ge 0\) if \(g^i \ge 0\).

  2. (ii)

    A constant \(n_s^{i,\infty }\) (depending on \(g^i\)) exists such that \(\theta ^i(\xi ,\varvec{z}) \rightarrow n_s^{i,\infty } L_s^i({\varvec{p}})\), as \(\xi \rightarrow (-1)^i\infty \), and the convergence is exponentially fast; in particular

    $$\begin{aligned} \Big \vert {\big \langle {\theta _s^i} \big \rangle }(\xi ) - n_s^{i,\infty } \Big \vert \le C\mathrm {e}^{-\alpha {\vert {\xi } \vert }}, \end{aligned}$$

    for some constants \(C>0\) and \(\alpha >0\). Moreover, \(n^{i,\infty } \ge 0\) if \(g^i \ge 0\).

  3. (iii)

    The Albedo operator, associating the inflow \(\theta ^i_\mathrm {in}\equiv g^i\) to the outflow \(\theta ^i_\mathrm {out}\),

    $$\begin{aligned} \mathcal {A}^i : X^i_\mathrm {in}\rightarrow X^i_\mathrm {out}, \quad \mathcal {A}^i \theta ^i_\mathrm {in}:= \theta ^i_\mathrm {out}, \end{aligned}$$

    is a compact linear operator.

We now come to the coupled Milne problem (A.2). Thanks to the Albedo operator just introduced, we can reformulate (A.2) as a Fredholm problem in \(X^1_\mathrm {in}\times X^2_\mathrm {in}\) for the unknown inflow data \((\theta ^1_\mathrm {in},\theta ^2_\mathrm {in})\), namely:

$$\begin{aligned} \begin{pmatrix} \theta ^1_\mathrm {in}\\ \theta ^2_\mathrm {in}\end{pmatrix} -\mathcal {K}\begin{pmatrix} \mathcal {A}^1 \theta ^1_\mathrm {in}\\ \mathcal {A}^2 \theta ^2_\mathrm {in}\end{pmatrix} = \begin{pmatrix} \varGamma ^1\\ \varGamma ^2 \end{pmatrix} \end{aligned}$$
(A.8)

where, recalling definition (4.23),

$$\begin{aligned} \mathcal {K}\begin{pmatrix} \theta ^1_\mathrm {out}\\ \theta ^2_\mathrm {out}\end{pmatrix} = \begin{pmatrix} \mathcal {K}^1\big (\theta ^1_\mathrm {out}, \theta ^2_\mathrm {out}\big ) \\ \mathcal {K}^2\big (\theta ^2_\mathrm {out}, \theta ^1_\mathrm {out}\big ) \end{pmatrix}, \end{aligned}$$

and the components of the non-homogeneous term are

$$\begin{aligned} \varGamma ^i = G^i_\mathrm {in}- \mathcal {K}^i\Big (G^i_\mathrm {out},G^j_\mathrm {out}\Big ). \end{aligned}$$
(A.9)

Let us now show that \(\mathcal {K}: X^1_\mathrm {out}\times X^2_\mathrm {out}\rightarrow X^1_\mathrm {in}\times X^2_\mathrm {in}\) is a linear, continuous operator. In fact, for \(\varvec{z}\in \varTheta ^i_\mathrm {in}\) we can write

$$\begin{aligned} \frac{\theta ^i_\mathrm {in}(\varvec{z})}{L^i(\varvec{z})} = \frac{\mathcal {K}^i(\theta ^i_{\mathrm {out}},\theta ^j_{\mathrm {out}})(\varvec{z})}{L^i(\varvec{z})} = R^i(\varvec{z}) \frac{\theta ^i_\mathrm {out}({\sim }\varvec{z})}{L^i(\varvec{z})} + T^j(\varvec{z}') \frac{ss'\theta ^j_\mathrm {out}(\varvec{z}')}{L^i(\varvec{z})}, \end{aligned}$$

where \(\varvec{z}' \in \varTheta ^j_\mathrm {out}\) is constrained to \(\varvec{z}\) by the conservation laws (3.26). Recalling definitions (4.3) and (A.1), and using (4.24) and the identity

$$\begin{aligned} \frac{\mathrm {e}^h}{(\mathrm {e}^h+1)^2} = \frac{\mathrm {e}^{-h}}{(\mathrm {e}^{-h}+1)^2}, \end{aligned}$$

it is not difficult to show that the following relation holds

$$\begin{aligned} \phi _1\big (A(n^i_s)\big )\, L^i_s({\varvec{p}}) = \phi _1\big (A(n^j_{s'})\big ) \, L_{s'}^j({\varvec{p}}'), \end{aligned}$$
(A.10)

for all \(\varvec{z}= ({\varvec{p}},s)\) and \(\varvec{z}' = ({\varvec{p}}',s')\) related as above. Then, the previous equality can be rewritten as

$$\begin{aligned} \frac{\theta ^i_\mathrm {in}(\varvec{z})}{L^i(\varvec{z})} = R^i(\varvec{z}) \frac{\theta ^i_\mathrm {out}({\sim }\varvec{z})}{L^i(\varvec{z})} + T^j(\varvec{z}') \frac{ss' c^i_s \theta ^j_\mathrm {out}(\varvec{z}')}{c^j_{s'} L^j(\varvec{z}')}, \end{aligned}$$

where

$$\begin{aligned} c^i(\varvec{z}) = c^i_s := \phi _1\big (A(n^i_s)\big ) \end{aligned}$$
(A.11)

are positive constants that only depend on s (and not on \({\varvec{p}}\)). Using Jensen inequality we can write

$$\begin{aligned} \Big \vert \frac{\theta ^i_\mathrm {in}(\varvec{z})}{L^i(\varvec{z})} \Big \vert ^2 \le R^i(\varvec{z}) \Big \vert \frac{\theta ^i_\mathrm {out}({\sim }\varvec{z})}{L^i(\varvec{z})} \Big \vert ^2 + T^j(\varvec{z}') \Big \vert \frac{c^i_s}{c^j_{s'}} \Big \vert ^2 \Big \vert \frac{\theta ^j_\mathrm {out}(\varvec{z}')}{ L^j(\varvec{z}')} \Big \vert ^2, \end{aligned}$$

(we adopt the redundant notation \({\vert {\cdot } \vert }^2\) for the square, just to improve readability), which shows the continuity of \(\mathcal {K}\), since the scattering coefficients are bounded by 1.

Hence, the composition of \(\mathcal {K}\), which is continuous, with the \(\mathcal {A}^i\)’s, which are compact, is a compact operator on \(X^1_\mathrm {in}\times X^2_\mathrm {in}\) and, therefore, (A.8) is a Fredholm equation with compact operator. The proof of Theorem 4.1 is thus reduced to a Fredholm alternative, which will be discussed in the next two steps.

Step 3: the homogeneous problem. We now consider the homogeneous version of the Milne problem (A.2), corresponding to \(G=0\):

$$\begin{aligned} \left\{ \begin{aligned}&\mu \frac{\partial \theta ^i}{\partial \xi } = L^i {\big \langle {\theta ^i} \big \rangle } - \theta ^i ,&(-1)^i \xi >0,\\&\theta ^i_\mathrm {in}- \mathcal {K}^i\big (\theta ^i_\mathrm {out},\theta ^j_\mathrm {out}\big ) = 0,&\xi =0. \end{aligned} \right. \end{aligned}$$
(A.12)

Let \((\theta ^1,\theta ^2)\) be a solution of such a problem in the space \(X^1\times X^2\). It is convenient to introduce the functions \((\psi ^1,\psi ^2)\) as follows:

$$\begin{aligned} \theta ^i(\xi ,\varvec{z}) = \sqrt{L^i(\varvec{z})} \, \psi ^i(\xi ,\varvec{z}), \quad i = 1,2, \end{aligned}$$

so that \(\psi ^i/\sqrt{L^i}\) is bounded and \(\psi ^i\) satisfies the equation

$$\begin{aligned} \mu \frac{\partial \psi ^i}{\partial \xi } = \sqrt{L^i} \, {\left\langle \sqrt{L^i} \, \psi ^i \right\rangle } - \psi ^i. \end{aligned}$$
(A.13)

It is immediate to verify that

$$\begin{aligned} \int _{\mathbb {R}^2} \left( \sqrt{L_s^i}\, \Big \langle \sqrt{L_s^i}\, \psi _s^i\Big \rangle - \psi _s^i\right) \psi ^i \, d{\varvec{p}}\le 0, \end{aligned}$$
(A.14)

and that the equality holds if and only if \(\psi _s^i\) is in the kernel of the collision operator, i.e. \(\psi _s^i = \sqrt{L_s^i} \, \gamma _s\), for some \(\gamma _s\) constant with respect to \({\varvec{p}}\). By multiplying by \(\psi ^i\) both sides of (A.13), integrating over \({\varvec{p}}\in \mathbb {R}^2\) we obtain

$$\begin{aligned} \frac{1}{2} \frac{\partial }{\partial \xi } \int _{\mathbb {R}^2} {\vert {\psi _s^i(\xi ,{\varvec{p}})} \vert }^2 \mu ({\varvec{p}})\, d{\varvec{p}}= \int _{\mathbb {R}^2} \left( \sqrt{L_s^i}\, \Big \langle \sqrt{L_s^i}\, \psi _s^i \Big \rangle - \psi _s^i\right) \psi _s^i \, d{\varvec{p}}\le 0. \end{aligned}$$
(A.15)

From (ii) of Step 2 we know that \(\psi _s^i(\xi ,{\varvec{p}})\), as \(\xi \rightarrow (-1)^i\infty \), tends to a function of the form \(\sqrt{L_s^i({\varvec{p}})}\, \gamma _s^i(\xi )\) and, therefore, by integrating the previous inequality over \(\xi \in (-1)^i[0,+\infty )\) and recalling that \(\mu \) is an odd function of \({\varvec{p}}\), we obtain

$$\begin{aligned} \int _{\mathbb {R}^2} {\vert {\psi _s^1(0,{\varvec{p}})} \vert }^2 \mu ({\varvec{p}}) \, d{\varvec{p}}\le 0 \le \int _{\mathbb {R}^2} {\vert {\psi _s^2(0,{\varvec{p}})} \vert }^2 \mu ({\varvec{p}}) \, d{\varvec{p}}. \end{aligned}$$
(A.16)

On the other hand, \((\theta ^1,\theta ^2)\) satisfy the homogeneous KTC

$$\begin{aligned} \theta ^i(0,\varvec{z}) = R^i(\varvec{z}) \theta ^i(0,{\sim }\varvec{z}) + T^j(\varvec{z}') ss' \theta ^j(0,\varvec{z}') \end{aligned}$$

and, correspondingly, \((\psi ^1,\psi ^2)\) satisfy

$$\begin{aligned} \sqrt{L^i(\varvec{z})}\, \psi ^i(0,\varvec{z}) = R^i(\varvec{z}) \sqrt{L^i(\varvec{z})}\, \psi ^i(0,{\sim }\varvec{z}) + T^j(\varvec{z}') ss' \sqrt{L^j(\varvec{z}')}\, \psi ^j(0,\varvec{z}'), \end{aligned}$$

where \(\varvec{z}\in \varTheta ^i_\mathrm {in}\) and \(\varvec{z}' \in \varTheta ^j_\mathrm {out}\) are constrained by the conservation of energy (3.26). By using (A.10) and (A.11), we obtain

$$\begin{aligned} \sqrt{c^i(\varvec{z})}\, \psi ^i(0,\varvec{z}) = R^i(\varvec{z}) \sqrt{c^i(\varvec{z})}\, \psi ^i(0,{\sim }\varvec{z}) + T^j(\varvec{z}') ss' \sqrt{c^j(\varvec{z}')}\, \psi ^j(0,\varvec{z}'), \end{aligned}$$
(A.17)

and then, using Jensen inequality,

$$\begin{aligned} c^i(\varvec{z}) {\vert {\psi ^i(0,\varvec{z})} \vert }^2 \le R^i(\varvec{z}) c^i(\varvec{z}) {\vert {\psi ^i(0,{\sim }\varvec{z})} \vert }^2 + T^j(\varvec{z}') c^j(\varvec{z}') {\vert {\psi ^j(0,\varvec{z}')} \vert }^2 \end{aligned}$$

or, equivalently,

$$\begin{aligned}&c^i(\varvec{z}) \left( {\vert {\psi ^i(0,\varvec{z})} \vert }^2- {\vert {\psi ^i(0,{\sim }\varvec{z})} \vert }^2\right) \\&\quad \le - T^i(\varvec{z}) c^i(\varvec{z}) {\vert {\psi ^i(0,{\sim }\varvec{z})} \vert }^2 + T^j(\varvec{z}') c^j(\varvec{z}'){\vert {\psi ^j(0,\varvec{z}')} \vert }^2. \end{aligned}$$

We now multiply both sides by \(\mu (\varvec{z})\), \(\varvec{z}\in \varTheta ^i_\mathrm {in}\), which is negative (or zero) for \(i = 1\) and positive (or zero) for \(i = 2\), so that

$$\begin{aligned} c^1(\varvec{z}) \left( {\vert {\psi ^1(0,\varvec{z})} \vert }^2- {\vert {\psi ^1(0,{\sim }\varvec{z})} \vert }^2\right) \mu (\varvec{z})\ge & {} - T^1(\varvec{z}) c^1(\varvec{z}) {\vert {\psi ^1(0,{\sim }\varvec{z})} \vert }^2\mu (\varvec{z})\\&+ T^2(\varvec{z}') c^2(\varvec{z}'){\vert {\psi ^2(0,\varvec{z}')} \vert }^2\mu (\varvec{z}),\\ c^2(\varvec{z}) \left( {\vert {\psi ^2(0,\varvec{z})} \vert }^2- {\vert {\psi ^2(0,{\sim }\varvec{z})} \vert }^2\right) \mu (\varvec{z})\le & {} - T^2(\varvec{z}) c^2(\varvec{z}) {\vert {\psi ^2(0,{\sim }\varvec{z})} \vert }^2\mu (\varvec{z})\\&+ T^1(\varvec{z}') c^1(\varvec{z}'){\vert {\psi ^1(0,\varvec{z}')} \vert }^2\mu (\varvec{z}). \end{aligned}$$

If we now integrate the first inequality over \(\varvec{z}\in \varTheta ^1_\mathrm {in}\), and the second one over \(\varvec{z}\in \varTheta ^2_\mathrm {in}\), by following the same passages as in the proof of Proposition 3.1 we arrive at

$$\begin{aligned} \int _\varTheta c^1(\varvec{z}) {\vert {\psi ^1(0,\varvec{z})} \vert }^2 \mu (\varvec{z}) d\varvec{z}\ge & {} \int _{\varTheta ^1_\mathrm {out}} T^1(\varvec{z}) c^1(\varvec{z}) {\vert {\psi ^1(0,\varvec{z})} \vert }^2 \mu (\varvec{z})\,d\varvec{z}\\&+ \int _{\varTheta ^2_\mathrm {out}} T^2(\varvec{z}') c^2(\varvec{z}') {\vert {\psi ^2(0,\varvec{z}')} \vert }^2 \mu (\varvec{z}')\,d\varvec{z}'\\ \int _\varTheta c^2(\varvec{z}) {\vert {\psi ^2(0,\varvec{z})} \vert }^2 \mu (\varvec{z}) d\varvec{z}\le & {} \int _{\varTheta ^2_\mathrm {out}} T^2(\varvec{z}') c^2(\varvec{z}') {\vert {\psi ^2(0,\varvec{z}')} \vert }^2 \mu (\varvec{z}')\,d\varvec{z}'\\&+ \int _{\varTheta ^1_\mathrm {out}} T^1(\varvec{z}) c^1(\varvec{z}) {\vert {\theta ^1(0,\varvec{z})} \vert }^2 \mu (\varvec{z})\,d\varvec{z}\end{aligned}$$

(where in the last integral of both inequalities, \(\varvec{z}\) and \(\varvec{z}'\) are constrained by \(E(\varvec{z}) = E(\varvec{z}') + {\delta V}\)), which immediately leads to

$$\begin{aligned} \int _\varTheta c^1(\varvec{z}) {\vert {\psi ^1(0,\varvec{z})} \vert }^2 \mu (\varvec{z}) d\varvec{z}\ge \int _\varTheta c^2(\varvec{z}) {\vert {\psi ^2(0,\varvec{z})} \vert }^2 \mu (\varvec{z}) d\varvec{z}. \end{aligned}$$
(A.18)

If we now come back to (A.16), multiply the right and the left sides by the positive constants \(c^1(\varvec{z}) = c^1_s\) and \(c^2(\varvec{z}) = c^2_s\), respectively, and sum up with respect to s, we obtain

$$\begin{aligned} \int _\varTheta c^1(\varvec{z}) {\vert {\psi ^1(0,\varvec{z})} \vert }^2 \mu (\varvec{z}) d\varvec{z}\le 0 \le \int _\varTheta c^2(\varvec{z}) {\vert {\psi ^2(0,\varvec{z})} \vert }^2 \mu (\varvec{z}) d\varvec{z}. \end{aligned}$$
(A.19)

By comparing (A.18) with (A.19) we see that, necessarily,

$$\begin{aligned} \int _\varTheta c^1(\varvec{z}) {\vert {\psi ^i(0,\varvec{z})} \vert }^2 \mu (\varvec{z}) d\varvec{z}= 0, \qquad i = 1,2. \end{aligned}$$

Multiplying (A.15) by \(c^i(\varvec{z}) = c^i_s\), summing up with respect to s and integrating with respect to \(\xi \) yields, therefore,

$$\begin{aligned} \int _0^{(-1)^i\infty } \int _\varTheta c^i \left( \sqrt{L^i}\, \Big \langle \sqrt{L^i}\, \psi ^i \Big \rangle - \psi ^i\right) \psi ^i \, d\varvec{z}\,d\xi = 0. \end{aligned}$$

Since \(c^i\) are positive constants that only depend on s, and the integrals with respect to \({\varvec{p}}\) are definite in sign (see (A.14)), this implies that

$$\begin{aligned} \int _{\mathbb {R}^2} \left( \sqrt{L_s^i}\, \Big \langle \sqrt{L_s^i}\, \psi _s^i \Big \rangle - \psi _s^i\right) \psi _s^i \,d{\varvec{p}}= 0 \end{aligned}$$

for all \(\xi \in (-1)^i[0,+\infty )\). This equality can only hold when \(\psi ^i(\xi ,\cdot )\) is in the kernel of the collision operator, which implies that \(\theta ^i(\xi ,{\varvec{p}})\) is necessarily of the form

$$\begin{aligned} \theta ^i(\xi ,\varvec{z}) = \theta _s^i(\xi ,{\varvec{p}}) = L_s^i({\varvec{p}}) \gamma _s^i(\xi ). \end{aligned}$$

Finally, the substitution of this expression in the first of equations (A.12) immediately yields that \(\gamma _s^i\) is constant, so that

$$\begin{aligned} \theta _s^i(\xi ,{\varvec{p}}) = L_s^i({\varvec{p}}) \gamma _s^i. \end{aligned}$$
(A.20)

Substituting (A.20) in the second of equations (A.12) and using (A.10) leads to the following necessary and sufficient condition for (A.20) to be solution of the homogeneous Milne problem (A.12):

$$\begin{aligned} c_{s'}^j\gamma _s^i = ss'c_s^i \gamma _{s'}^j \end{aligned}$$
(A.21)

where \(c^i_s\) is defined by (A.11), and ij and \(s,s'\) are related as usual. Recalling that the conservation of energy is satisfied by three couples \((s,s')\) if \({\delta V}\not = 0\) and just by two couples in the case if \({\delta V}= 0\) (see Remark 3.1), we notice that (A.21) is a rank-3 condition if \({\delta V}\not = 0\) and a rank-2 condition if \({\delta V}= 0\).

Step 4: the inhomogeneous problem. Let us finally return to the complete, inhomogeneous problem (A.2). Let \((\theta ^1,\theta ^2)\) be a bounded solution of (A.2). The integration in \({\varvec{p}}\) of the first equation in (A.2) yields

$$\begin{aligned} \frac{d}{d\xi } {\left\langle \mu \theta _s^i \right\rangle } = 0, \end{aligned}$$

and the integration in \({\varvec{p}}\) after multiplication by \(\mu \) yields

$$\begin{aligned} \frac{d}{d\xi } {\left\langle \mu ^2 \theta _s^i \right\rangle } = - {\left\langle \mu \theta _s^i \right\rangle }. \end{aligned}$$

Hence, \({\left\langle \mu \theta _s^i \right\rangle }\) is a constant, and this constant must be zero, otherwise \({\vert {{\left\langle \mu ^2 \theta _s^i \right\rangle }} \vert }\) would grow linearly with \(\xi \), in contradiction with the boundedness assumption. So we have

$$\begin{aligned} \int _{\mathbb {R}^2} \theta ^i_s(\xi ,{\varvec{p}})\,\mu ({\varvec{p}})\,d{\varvec{p}}= 0, \end{aligned}$$
(A.22)

for all \(\xi \in (-1)^i[0,+\infty )\), \(i =1,2\) and \(s = \pm 1\). If we now rewrite the boundary conditions as

$$\begin{aligned} \theta ^i_\mathrm {in}- G^i_\mathrm {in}- \mathcal {K}^i\Big (\theta ^i_\mathrm {out}- G^i_\mathrm {out},\theta ^j_\mathrm {out}- G^j_\mathrm {out}\Big ) = 0, \end{aligned}$$

then, from Proposition 3.1 (that applies also to the linear KTC), we have that the conservation of charge flux holds:

$$\begin{aligned} \int _\varTheta s(\theta ^1(0,\varvec{z})- G^1(\varvec{z})) \mu (z)\,d\varvec{z}= \int _\varTheta s(\theta ^2(0,\varvec{z})- G^2(\varvec{z})) \mu (\varvec{z})\,d\varvec{z}. \end{aligned}$$

But then, since (A.22) implies that the charge flux associated to \(\theta ^i\) vanishes, we obtain that the flux conservation for \(G^i\) must hold:

$$\begin{aligned} \int _\varTheta s\,G^1(\varvec{z}) \mu (\varvec{z})\,d\varvec{z}= \int _\varTheta s\, G^2(\varvec{z}) \mu (\varvec{z})\,d\varvec{z}. \end{aligned}$$
(A.23)

Equation (A.23) is therefore a necessary condition for the existence of a bounded solution to the Milne problem (A.2) or, equivalently, to the Fredholm problem (A.8) when \(\varGamma ^i\) has the form (A.9). Using (4.14), it is immediate to verify that the \(G^i\)’s, given by (4.18), satisfy this condition if and only if (4.30) holds.

Now, from Step 3 we know that the kernel of the Fredholm operator at the left-hand side of (A.8) is spanned by the functions of the form \(L^i_s({\varvec{p}})\gamma ^i_s\), with \(\gamma ^i_s\) satisfying (A.21), and is therefore a subspace of \(X^1_\mathrm {in},\times X^2_\mathrm {in}\) of dimension d, where \(d = 1\) if \({\delta V}\not = 0\), and \(d = 2\) if \({\delta V}= 0\). Hence, the range of the Fredholm operator is a closed subspace of codimension d. But (A.23) also defines a subspace of codimension d, and then it describes the condition of existence of the solution to the Fredholm equation when \(\varGamma ^i\) is of the form (A.9). We conclude that (4.30) is a necessary and sufficient condition for the existence of a solution to the Fredholm equation (A.8) (and, therefore, to the Milne problem (A.2)), up to a solution of the associated homogeneous problem. This proves the first part of Theorem 4.1.

The second part of the theorem, that is the existence of the asymptotic densities \(n_s^{i,\infty }\) and the exponential estimate (4.32), follows from (ii) of Step 2. In fact, once the coupled Milne problem is solved, the inflow of each component \(\theta ^i_s\) is determined (up to the addition of a term of the form (A.20)–(A.21)), and point (ii) of Step 2 applies.Footnote 7

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Barletti, L., Negulescu, C. Quantum Transmission Conditions for Diffusive Transport in Graphene with Steep Potentials. J Stat Phys 171, 696–726 (2018). https://doi.org/10.1007/s10955-018-2032-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10955-018-2032-y

Keywords

Navigation