Acta Polytechnica https://doi.org/10.14311/AP.2022.62.0023 Acta Polytechnica 62(1):23–29, 2022 © 2022 The Author(s). Licensed under a CC-BY 4.0 licence Published by the Czech Technical University in Prague PHOTONIC GRAPHENE UNDER STRAIN WITH POSITION-DEPENDENT GAIN AND LOSS Miguel Castillo-Celeitaa, ∗, Alonso Contreras-Astorgab, David J. Fernández C.a a Cinvestav, Physics Department, P.O. Box. 14-740, 07000 Mexico City, Mexico b Cinvestav, CONACyT – Physics Department, P.O. Box. 14-740, 07000 Mexico City, Mexico ∗ corresponding author: mfcastillo@fis.cinvestav.mx Abstract. We work with photonic graphene lattices under strain with gain and loss, modeled by the Dirac equation with an imaginary mass term. To construct such Hamiltonians and their solutions, we use the free-particle Dirac equation and then a matrix approach of supersymmetric quantum mechanics to generate a new Hamiltonian with a magnetic vector potential and an imaginary position-dependent mass term. Then, we use a gauge transformation that maps our solutions to the final system, photonic graphene under strain with a position-dependent gain/loss term. We give explicit expressions for the guided modes. Keywords: Graphene, Dirac materials, photonic graphene, matrix supersymmetric, quantum mechan- ics. 1. Introduction Graphene is the last known carbon allotrope, it was isolated for the first time by Novoselov, Geim, et al. in 2004 [1]. This material consists of a two-dimensional hexagonal arrangement of carbon atoms. Graphene excels for its interesting properties, such as mechanical resistance, electrical conductivity, and optical opac- ity [2, 3]. The study of graphene has contributed to the development of different areas in physics, for example, in solid-state, graphene has prompted the discovery of other materials with similar characteristics, such as borophene and phosphorene. At low energy, the charge carriers in graphene behave like Dirac massless particles, and from this approach, graphene has al- lowed the verification of the Klein tunneling paradox as well as the quantum Hall effect. These phenomena have gained a special interest in particle physics and quantum mechanics [4]. Exploring graphene in an external constant mag- netic field has allowed identifying the discrete bound states in the material, the so-called Landau levels. Moreover, theoretical physicist have analyzed the be- havior of Dirac electrons in graphene under different magnetic field profiles as well. Supersymmetric quan- tum mechanics is a useful tool to find solutions of the Dirac equation under external magnetic fields [5–9]. Following this approach, a mechanical deformation in a graphene lattice is equivalent to introducing an external magnetic field [10, 11]. Graphene has its analog in photonics, called pho- tonic graphene. It is constructed through a two- dimensional photonic crystal with weakly coupled optical fibers in a three-dimensional setting [12–17]. Photonic graphene under strain is modeled through a deformation in the coupled optical fiber lattice [18– 21]. Compared with the conventional graphene Hamilto- nian, the photonic graphene Hamiltonian has an extra term that represents the gain/loss in the fibers. The literature on this topic always considers a constant gain/loss in space. With the previous motivation, we will apply supersymmetric quantum mechanics in a matrix approach (matrix SUSY-QM) to obtain solutions of the Dirac equation for strain photonic graphene with a position-dependent gain/loss. 2. Strain in photonic graphene The graphene structure consists of carbon atoms in a hexagonal arrangement similar to a honeycomb lat- tice. This structure can be described by two triangular sublattices of atoms, which are denoted as type A and type B. The base vectors to the unitary cell are given by a1 = a 2 ( √ 3, 3), a2 = a 2 (− √ 3, 3), (1) where a is the interatomic distance, for graphene a = 1.42 Å (see Figure 1a). The position of the atoms in the whole lattice can be defined by the set of vectors Rl = l1a1 + l2a2, with l1, l2 ∈ Z. An alternative description of graphene is through the first neighbors, which are connected by the vectors δn δ1 = a 2 ( √ 3, 1), δ2 = a 2 (− √ 3, 1), δ3 = a(0, −1). (2) A reciprocal lattice can be defined in the momentum space, which is also hexagonal, as shown in Figure 1b. It is rotated 90◦ with respect to the original carbon network. A hexagon in the reciprocal lattice is recog- nized as the first Brillouin zone. In this zone, there are only two inequivalent points, K± = (± 4π3√3a, 0). 23 https://doi.org/10.14311/AP.2022.62.0023 https://creativecommons.org/licenses/by/4.0/ https://www.cvut.cz/en M. Castillo-Celeita, A. Contreras-Astorga, D. J. Fernández C. Acta Polytechnica A B a1a2 δ1δ2 δ3 (a). b1 b2 K+K- (b). Figure 1. (A) Hexagonal graphene lattice. The lat- tice is constructed by type A and type B atoms, in this case, a1 and a2 correspond to the lattice unitary vectors, and δn are the vectors that connect the atoms A(B) with the nearest neighbors. (B) Reciprocal lat- tice, which is characterized by the b1,2 vectors and K± correspond to the two possible inequivalent points in the lattice. All subsequent corners are determined from either K+ or K− plus integer multiples of the vectors b1 = 2π 3a ( √ 3, 1), b2 = 2π 3a (− √ 3, 1). (3) Vectors ai and bj fulfill the condition ai · bj = 2πδij . 2.1. Tight-binding model The tight-binding Hamiltonian describes the hopping of an electron from an atom A (B) to an atom B (A) H = −t ∑ Ri 3∑ n=1 (|ARi ⟩ ⟨BRi +δn | + |BRi +δn ⟩ ⟨ARi |), (4) where t ≈ 3 eV is called the hopping integral, Ri runs over all sites in the sublattice A, thus |ARi ⟩ is a state vector in these sites, the same applies to B and |BRi+δn ⟩, recall that δn connects the atoms of the sublattice A(B) with its nearest neighbors in the sublattice B(A). The translational symmetry suggests the use of Bloch states |ΨBloch⟩ = 1 √ Nc ∑ Rj (eik·Rj ψA(k) |ARj ⟩ + eik·(Rj +δ3)ψB (k) |BRj +δ3 ⟩), (5) where Nc is the number of the unitary cell [22]. Then H |Ψ⟩ = E |Ψ⟩ becomes a matrix problem 0 −t 3∑ n=1 e−ik·δn −t 3∑ n=1 eik·δn 0 ( ψA ψB ) = E ( ψA ψB ) , (6) with ψA ≡ ψA(k) and ψB ≡ ψB (k) and the energy term given by E± = ± ∣∣∣∣∣t 3∑ n=1 e−ik·δn ∣∣∣∣∣ = ±t √ 3 + 2 cos( √ 3kxa) + 4 cos( √ 3kx a 2 ) cos(3ky a 2 ). To obtain an effective Hamiltonian at low energy, we can consider the Taylor series around the Dirac points H(k = K± +q) ≈ q · ∇kH|K± . Note that E(K±) = 0, as a consequence, at these points, the valence and conduction bands are connected. The above calculus leads to the analog of the Dirac-Weyl equation HϱΨ = ℏv0(ϱσ1qx + σ2qy )Ψ = EΨ, (7) where ϱ = ±1 correspond to the K± valleys, v0 is called the Fermi velocity, in graphene, v0 = 3ta/2ℏ ≈ c/300, with c being the velocity of light, σi are the Pauli matrices σ1 = ( 0 1 1 0 ) , σ2 = ( 0 −i i 0 ) , σ3 = ( 1 0 0 −1 ) , (8) and Ψ is a bi-spinor. The matrix nature of this equa- tion is related to the sublattices A and B, this degree of freedom is called pseudo-spin. Notice that at low energies, the dispersion relation is linear, given by E±(q) = ±ℏv0|q|, then, the Dirac cones are connect- ing at E± = 0, as expected for particles without mass [23]. 2.2. Uniform strain The photonic analog of a graphene lattice is built with weakly coupled optical fibers. This kind of photonic system is described by the same tight-binding Hamil- tonian in graphene with an additional term γA/B ; that represents the gain and loss in the optical fibers in the position A/B, this new term produces an attenuation or amplification in the optical modes. If we consider uniform strain in the lattices, which is represented by a strain tensor u = ( u11 0 0 u22 ) , (9) the Fermi velocity is modified in the following form vij = v0(1 + (1 − β)uij ). (10) The hopping integrals are modified with a little per- turbation t → tn, that, considering the changes in the orbitals by the modification of the carbon distances tn ≈ t ( 1 − β a2 δn · u · δn ) , (11) 24 vol. 62 no. 1/2022 Photonic graphene under strain with position-dependent . . . where β = − ∂ ln t ∂ ln a (12) is the Grüneisen parameter that depends on the model; for graphene, β is between 2 and 3 [10] (see also [24, 25]). In photonic graphene, a is the distance between adjacent waveguides. The Hamiltonian of a photonic graphene with a uniform strain reads as H = γA ∑ Ri |ARi ⟩ ⟨ARi | + γB ∑ Ri |BRi +δn ⟩ ⟨BRi +δn | − ∑ Ri 3∑ n=1 tn(|ARi ⟩ ⟨BRi +δn | + |BRi +δn ⟩ ⟨ARi |). (13) The deformation of the lattice produces a shift of the Dirac points KD± ≈ (1 − u) · K± ± A, where A = (Ax, Ay ) Ax = β 2a (u11 − u22), Ay = − β 2a (2u12). (14) Using the Bloch solution, the Hamiltonian under strain takes the form: H = γA − 3∑ n=1 tne −ik·(1−u)·δn − 3∑ n=1 tne ik·(1−u)·δn γB , (15) under the assumption |u·δn| ≪ a. In this work, we will assume that γA = iγ and γB = γ∗A, then, for positive γ, the waveguides in the sublattice A (B) present the energy gain (loss), as in the arrangements proposed in [14]. Expanding this Hamiltonian around the Dirac points, through the substitution k = KD± + q, one arrives at a Dirac Hamiltonian analog with minimal coupling H = v0σ · (1 + u − βu)q + iγσ3. (16) Comparing with (7), the effect of strain is equiva- lent, to consider magnetic-like field modeled through a pseudo-magnetic vector potential A. The last term represents a gain/loss balance in sublattices A/B. In photonic graphene, strain could be generated by de- formations in the geometry of the optical-fiber lattice. 2.3. Non-uniform strain For non-uniform strain, the deformation matrix de- pends of the position, u → u(r). Thus, the expression for the Hamiltonian becomes H = −iσi √ vij∂j √ vij + v0σiAi + iγσ3, (17) considering a strain tensor of the form u = ( u11(x) 0 0 u22(y) ) , (18) and equations (10) and (14), still apply. We can also write the strain Hamiltonian as H(x,y) = − iσ1 √ v11∂x √ v11 − iσ2 √ v22∂y √ v22 + σ1 v0β 2a (u11(x) − u22(y)) + iγσ3, (19) where v11 = v11(x), v22 = v22(y). We can relate the eigenvalue equation of this Hamil- tonian, HΨ = EΨ, with a strain-free one using the following transformation. First, we define the coordi- nates r = ∫ v0 v11(x) dx, s = ∫ v0 v22(y) dy, (20) and the operator G(x,y) = √ v11v22 v0 exp ( iv0β 2a ∫ x 0 u11(q) v11(q) dq ) , (21) then, H will be related with a flat Fermi velocity Hamiltonian H0 as H(x,y) = G−1(x,y)H0(r(x),s(y))G(x,y), (22) where H0Φ = ( −iv0σ1∂r − iv0σ2∂s − v0β 2a u22 + iγσ3 ) Φ, (23) and u22 = u22(y(r,s)). The solutions are mapped as Ψ(x,y) = G−1(x,y)Φ(r(x),s(y)). (24) The energy spectrum is the same for both Hamiltoni- ans [18, 19, 26]. 3. Supersymmetric quantum mechanics: matrix approach Supersymmetric quantum mechanics (SUSY-QM) is a method that relates two Schrödinger Hamiltonians through an intertwining operator [27, 28]. Another approach is the matrix SUSY-QM, which intertwines two Dirac Hamiltonians H0, H1 by a matrix operator L. In this work, we use the latter to construct an appropriate Hamiltonian H1 that will be linked via the operator G introduced in (21) to a photonic graphene system under strain. For the sake of completeness we will give a brief review of matrix SUSY-QM (more details can be found in [29]). We start by proposing the following intertwining relation: L1H0 = H1L1, (25) where the Dirac Hamiltonians are given by H0 = −iσ2∂s + V0(s), H1 = −iσ2∂s + V1(s), (26) and the intertwining operator is L1 = ∂s − UsU−1, (27) with U being a matrix function called seed or trans- formation matrix, the subindex in Us represent the derivative respect to s, and U must satisfy H0U = UΛ. Let us write U in a general form and Λ as a diagonal matrix U = ( u11 u12 u21 u22 ) , Λ = ( λ1 0 0 λ2 ) . (28) 25 M. Castillo-Celeita, A. Contreras-Astorga, D. J. Fernández C. Acta Polytechnica From the intertwining relation and the given defini- tions, the potential V1 can be written in terms of the potential V0 and the transformation matrix as V1 = V0 + i[UsU−1,σ2]. (29) Solutions of the Dirac equation H0ξ = Eξ can be mapped onto solutions of H1Φ = EΦ using the inter- twining operator as Φ ∝ L1ξ. There are some extra solutions, usually referred to missing states. They can be obtained from each column of (UT )−1, named Φλj , j = 1, 2, which satisfy H1Φλj = λj Φλj . If the vectors Φλj fulfill the boundary conditions of the prob- lem, λj must be included in the spectrum of H1. As a summary, with this technique, we start from H0, its eigenspinors and spectrum, then we construct H1, obtain the solutions of the corresponding Dirac equa- tion and the spectrum. Now, let us mention that it is possible to iterate this technique. The main advantage comes from the modification of the spectrum, since with each iteration, we can add more energy levels. The second-order matrix SUSY-QM can be reached through a second intertwining relation L2H1 = H2L2, (30) which is similar to (25). The intertwining operator now takes the form L2 = ∂s − (U2)sU −12 . (31) The operator L1 is used to determine the transforma- tion matrix of the second iteration, U2 = L1U2, where U2 fulfills the relation H0U2 = U2Λ2. In this case, Λ2 is an Hermitian matrix that we choose diagonal once again, Λ2 = ( λ̃1 0 0 λ̃2 ) , U2 = ( w11 w12 w21 w22 ) . (32) The elements of Λ2 are such that (λ̃1, λ̃2) ̸= (λ1,λ2). Therefore, the second order potential is given by V2 = V1 + i [ (U2)sU −12 ,σ2 ] . (33) The solutions of H2χ = Eχ are obtained from the eigenspinors of H1 as χ ∝ L2Φ. The second-order matrix SUSY-QM generates, in principle, two sets of eigenspinors that correspond to the columns of the matrices (U T2 )−1 and L2(UT )−1. 4. Photonic graphene under strain and position-dependent gain and loss In this section, we start from the auxiliary Dirac equation of a free particle with imaginary mass, and using a matrix SUSY-QM and a gauge transformation G, we obtain a photonic graphene model with strain and position dependent gain/loss. We show that we can iterate the technique to add more propagations modes. Figure 2. Graph of the functions v0kr + K(s) (line blue) and the gain/loss term γ − Γ(s) (dashed red line) for ϵ = 1.5, kr = π, γ = 1, v0 = 1.0. Notice that γ − Γ(s) coincides asymptotically with γ. 4.1. Photonic Graphene with a single mode Let us start from the free particle Dirac equation where we included a purely imaginary mass term: H0Φ =(−iv0σ1∂r − iv0σ2∂s + iγσ3)Φ. (34) Considering Φ(r,s) = exp(ikrr)(ϕA(s),ϕB (s))T , the Hamiltonian can be written as H0(r,s) = −iv0σ2∂s + V0, (35) where V0 = v0krσ1 + iγσ3. Now, we use the matrix SUSY-QM to construct a new system. A convenient selection of the Λ elements is λ1 = ϵ = −λ2. We build the transformation matrix U with the entries u21 = u∗22 = cosh(κs) + i sinh(κs), the corresponding momentum in s is given by κ = √ k2r − (γ2 + ϵ2)/v20 . The other two components are found through the equation: u1j = v0 (λj − iγ) (−u′2j + kru2j ), j = 1, 2. (36) From (29) we obtain V1 as V1 = V0 + σ1K(s) − iσ3Γ(s), (37) where Γ(s), K(s) are given by Γ = 2γ + 2ϵ (κ(γ sinh(2κs) + ϵ) − γkr cosh(2κs)) κ(γ − ϵ sinh(2κs)) + krϵ cosh(2κs) , K = 2v0krϵ (kr cosh(2κs) − κ sinh(2κs)) κ(γ − ϵ sinh(2κs)) + krϵ cosh(2κs) − 2v0kr. Figure 2 shows a plot of the functions v0kr + K(s) and γ − Γ(s). The new Hamiltonian takes the form H1(r, s) = −iσ2v0∂s + σ1(−iv0∂r + K) + iσ3(γ − Γ). (38) This system supports two single bound states. They are the columns of the matrix (UT )−1 = (Φϵ, Φ−ϵ). The eigenvector associated with ϵ is given by Φϵ(r,s) = eikr r 2 − (γ2+ϵ2)(cosh(κs)+i sinh(κs))v0κ(γ−ϵ sinh(2κs))+v0kr ϵ cosh(2κs) (γ−iϵ)((κ+ikr ) cosh(κs)−(kr +iκ) sinh(κs)) κ(γ−ϵ sinh(2κs))+kr ϵ cosh(2κs) . (39) 26 vol. 62 no. 1/2022 Photonic graphene under strain with position-dependent . . . Our next step is to apply the gauge transformation defined in (20)-(22). The strain and Fermi velocity tensors that we consider are u = ( 0 0 0 − 2aK(y)) β ) , v = v0 ( 1 1 1 1 − (1 − β) 2aK(y) β ) , (40) see (18). The change of variables in (20) becomes r = x, s = ∫ 1 1 − (1 − β) 2aK(y) β dy, (41) and the operator G(x,y) = √ 1 − (1 − β) 2aK(y) β . This choice leads to the following Hamiltonian H1(x,y) = − iv0σ1∂x − iσ2 √ v22(y)∂y √ v22(y) − σ1 v0β 2a u22(y) + i(γ − Γ(y))σ3. (42) Bounded eigenstates of H1 can be found as Ψϵ(x,y) = G−1(x,y)Φϵ(r(x),s(y)). In this system, there is a sin- gle mode in the upper Dirac cone and another in the bottom cone. The strain generates an analog of a magnetic field perpendicular to the graphene layer B⃗(y) = (β/2a)∂y u22ẑ. Since we are working with a photonic graphene, such a pseudo-magnetic field affects light. Moreover, the term iΓ(y)σ3 indicates a position de- pendent gain/loss in the optical fibers of the sublattice A/B. Figure 3a shows the square modulus of each com- ponent of Φϵ = (ϕϵA,ϕϵB )T (shadowed curves) and the intensity |ϕϵA|2 + |ϕϵB |2 (red curve). Figure 3b shows the same for the mode Ψϵ. 4.2. Photonic Graphene with two modes In this subsection, we use two iterations of the ma- trix SUSY-QM, starting again from the free-particle Hamiltonian. Let us choose an initial system with zero gain/loss (γ = 0), which is a massless fermion in graphene. In the first matrix SUSY-QM step we use the same transformation matrix U as in the ex- ample above. The first matrix SUSY-QM partner Hamiltonian has the form H1 = −iσ2v0∂s + σ1K(s) + iσ3Γ(s), (43) where K(s) = krv0, Γ(s) = 2ϵv0κ κv0 sinh(2κs) − krv0 cosh(2κs) , (44) with κ = √ (krv0)2 − ϵ2/v0. As a result of the first matrix SUSY-QM step, it is generated a position de- pendent gain/loss term Γ(s). The iteration of the method requires to define the second diagonal ma- trix Λ2, with λ̃1 = −λ̃2 = ϵ1 ≠ ϵ, and the sec- ond transformation matrix U2. For this example, we (a). (b). Figure 3. (A) Plot of the individual intensities |ϕϵA|2 (gray curve) and |ϕϵB |2 (blue curve) and the total intensity |ϕϵA|2 + |ϕϵB |2 (red line). (B) Analog of the (A) plot for the solution Ψϵ of the Hamiltonian under strain. The parameters in this case are: kr = π, ϵ = 1.5, a = 1.0, β = 0.8, γ = 1, v0 = 1.0. Figure 4. Graph of the function v0kr + K2(s) (black line), and the gain/loss function Γ(s) + Γ2(s) (red dashed line), for ϵ = 1.5, ϵ1 = 2.0, kr = π, γ = 0, v0 = 1.0. choose w21 = cosh(κ2s) and w22 = cosh(κ2s), where κ2 = √ (krv0)2 − ϵ21/v0. The other two components can be found through the equation w1j = v0 λ̃j (−w′2j + krw2j ), j = 1, 2. (45) The potential V2 can be calculated from (33), V2 = V1+σ1K2(s)+iσ3Γ2(s) = V0+σ1K2+iσ3(Γ+Γ2). The functions v0kr + K2(s) and Γ(s) + Γ2(s) are shown in Figure 4. It is important to highlight that the gain/loss term remains a pure imaginary quantity. 27 M. Castillo-Celeita, A. Contreras-Astorga, D. J. Fernández C. Acta Polytechnica (a). (b). Figure 5. (A) Intensity of the superposition |Φ̄(s, z)|2 propagating in the z-axis. (B) Intensity of the super- position |Ψ̄(y, z)|2. The values of the parameters taken are ϵ = 1.5, ϵ1 = 2.0, kr = π, γ = 0, β = 0.8, a = 1.0, v0 = 1.0. The second matrix SUSY-QM step introduces two new sets of eigenmodes. They can be extracted from the columns of the matrix (U T2 )−1 = (χϵ1,χ−ϵ1 ). The eigenmodes added in the first step are mapped as χ±ϵ = L2Φ±ϵ. Similar to the previous example, it is possible to perform the gauge transformation (20)- (22). Then, in the system under strain, the modes become Ψ±ϵ1 (x,y) = G −1(x,y)χ±ϵ1 (r(x),s(y)), Ψ±ϵ(x,y) = G−1(x,y)χ±ϵ(r(x),s(y)). Therefore, in this new optical system, two guided modes are created in the upper Dirac cone and two more in the bottom Dirac cone. Finally, let us mention that we can have superpositions of the introduced modes and let them propagate along the z-axis inside the photonic graphene. For example, in the flat Fermi velocity system (before the gauge transformation), Φ̄(s,z) = A1e−iϵz Φϵ(s) + A2e−iϵ1z Φϵ1 (s), (46) becomes Ψ̄(y,z) = A1e−iϵz Ψϵ(y) + A2e−iϵ1z Ψϵ1 (y), (47) in the photonic graphene system under strain with the position dependent gain/loss balance. Figure 5a shows the propagation along z-axis of the intensity |Φ̄(s,z)|2, while Figure 5b shows |Ψ̄(s,z)|2. 5. Summary This article shows a natural way to construct Hamil- tonians associated with a photonic graphene under strain with a position-dependent gain/loss balance. The main tools that we use are a matrix approach to supersymmetric quantum mechanics and a gauge transformation. With a correct choice of a transforma- tion matrix U, it is possible to add a bound state to the free-particle Hamiltonian using the matrix SUSY-QM, but the Dirac equation will have two new terms in the potential, V1 = V0 + σ1K(s) − iσ3Γ(s). The function K could be associated with a magnetic vector poten- tial, but the function iΓ is related to an imaginary mass term, which is difficult to interpret or realize in a solid-state graphene. The gauge transformation G maps solutions from the flat Fermi velocity system of the previous step to a graphene system under strain. At this point, it becomes relevant to work with the photonic graphene. The magnetic vector potential translates into deformations of the lattice of optical fibers, while the iΓ function indicates the gain/loss of the fibers in the sublattice A/B. We end with the Hamiltonian of photonic graphene with a single mode. This mode is confined by the strain and the position- dependent gain/loss balance. Finally, we show that the technique can be iterated, to have two or more modes in the photonic graphene. Acknowledgements The authors acknowledge the support of Conacyt, grant FORDECYT-PRONACES/61533/2020. M. C-C. acknowl- edges as well the Conacyt fellowship 301117. References [1] K. S. Novoselov, A. K. Geim, S. V. Morozov, et al. Electric field effect in atomically thin carbon films. Science 306(5696):666–669, 2004. https://doi.org/10.1126/science.1102896. [2] R. C. Andrew, R. E. Mapasha, A. M. Ukpong, N. Chetty. Mechanical properties of graphene and boronitrene. Physical Review B 85(12):125428, 2012. https://doi.org/10.1103/PhysRevB.85.125428. [3] S.-E. Zhu, S. Yuan, G. C. A. M. Janssen. Optical transmittance of multilayer graphene. EPL (Europhysics Letters) 108(1):17007, 2014. https://doi.org/10.1209/0295-5075/108/17007. [4] A. K. Geim, K. S. Novoselov. The rise of graphene. In Nanoscience and technology: a collection of reviews from nature journals, pp. 11–19. 2009. https://doi.org/10.1142/9789814287005_0002. [5] Ş. Kuru, J. Negro, L. M. Nieto. Exact analytic solutions for a Dirac electron moving in graphene under magnetic fields. Journal of Physics: Condensed Matter 21(45):455305, 2009. https://doi.org/10.1088/0953-8984/21/45/455305. 28 https://doi.org/10.1126/science.1102896 https://doi.org/10.1103/PhysRevB.85.125428 https://doi.org/10.1209/0295-5075/108/17007 https://doi.org/10.1142/9789814287005_0002 https://doi.org/10.1088/0953-8984/21/45/455305 vol. 62 no. 1/2022 Photonic graphene under strain with position-dependent . . . [6] B. Midya, D. J. Fernández C. Dirac electron in graphene under supersymmetry generated magnetic fields. Journal of Physics A: Mathematical and Theoretical 47(28):285302, 2014. https://doi.org/10.1088/1751-8113/47/28/285302. [7] A. Contreras-Astorga, A. Schulze-Halberg. The confluent supersymmetry algorithm for Dirac equations with pseudoscalar potentials. Journal of Mathematical Physics 55(10):103506, 2014. https://doi.org/10.1063/1.4898184. [8] M. Castillo-Celeita, D. J. Fernández C. Dirac electron in graphene with magnetic fields arising from first-order intertwining operators. Journal of Physics A: Mathematical and Theoretical 53(3):035302, 2020. https://doi.org/10.1088/1751-8121/ab3f40. [9] A. Contreras-Astorga, F. Correa, V. Jakubský. Super-Klein tunneling of Dirac fermions through electrostatic gratings in graphene. Physical Review B 102(11):115429, 2020. https://doi.org/10.1103/PhysRevB.102.115429. [10] G. G. Naumis, S. Barraza-Lopez, M. Oliva-Leyva, H. Terrones. Electronic and optical properties of strained graphene and other strained 2D materials: a review. Reports on Progress in Physics 80(9):096501, 2017. https://doi.org/10.1088/1361-6633/aa74ef. [11] M. Oliva-Leyva, C. Wang. Theory for strained graphene beyond the Cauchy–Born rule. Physica Status Solidi (RRL)–Rapid Research Letters 12(9):1800237, 2018. https://doi.org/10.1002/pssr.201800237. [12] M. Polini, F. Guinea, M. Lewenstein, et al. Artificial honeycomb lattices for electrons, atoms and photons. Nature Nanotechnology 8(9):625–633, 2013. https://doi.org/10.1038/nnano.2013.161. [13] Y. Plotnik, M. C. Rechtsman, D. Song, et al. Observation of unconventional edge states in ‘photonic graphene’. Nature Materials 13(1):57–62, 2014. https://doi.org/10.1038/nmat3783. [14] H. Ramezani, T. Kottos, V. Kovanis, D. N. Christodoulides. Exceptional-point dynamics in photonic honeycomb lattices with PT symmetry. Physical Review A 85(1):013818, 2012. https://doi.org/10.1103/PhysRevA.85.013818. [15] G. G. Pyrialakos, N. S. Nye, N. V. Kantartzis, D. N. Christodoulides. Emergence of type-II Dirac points in graphynelike photonic lattices. Physical Review Letters 119(11):113901, 2017. https://doi.org/10.1103/PhysRevLett.119.113901. [16] S. Grosche, A. Szameit, M. Ornigotti. Spatial Goos-Hänchen shift in photonic graphene. Physical Review A 94(6):063831, 2016. https://doi.org/10.1103/PhysRevA.94.063831. [17] T. Ozawa, A. Amo, J. Bloch, I. Carusotto. Klein tunneling in driven-dissipative photonic graphene. Physical Review A 96(1):013813, 2017. https://doi.org/10.1103/PhysRevA.96.013813. [18] A. Szameit, M. C. Rechtsman, O. Bahat-Treidel, M. Segev. PT -symmetry in honeycomb photonic lattices. Physical Review A 84(2):021806, 2011. https://doi.org/10.1103/PhysRevA.84.021806. [19] H. Schomerus, N. Y. Halpern. Parity anomaly and Landau-level lasing in strained photonic honeycomb lattices. Physical Review Letters 110(1):013903, 2013. https://doi.org/10.1103/PhysRevLett.110.013903. [20] M. C. Rechtsman, J. M. Zeuner, A. Tünnermann, et al. Strain-induced pseudomagnetic field and photonic Landau levels in dielectric structures. Nature Photonics 7(2):153–158, 2013. https://doi.org/10.1038/nphoton.2012.302. [21] D. A. Gradinar, M. Mucha-Kruczyński, H. Schomerus, V. I. Fal’ko. Transport signatures of pseudomagnetic Landau levels in strained graphene ribbons. Physical Review Letters 110(26):266801, 2013. https://doi.org/10.1103/PhysRevLett.110.266801. [22] C. Bena, G. Montambaux. Remarks on the tight-binding model of graphene. New Journal of Physics 11(9):095003, 2009. https://doi.org/10.1088/1367-2630/11/9/095003. [23] P. Dietl, F. Piéchon, G. Montambaux. New magnetic field dependence of Landau levels in a graphenelike structure. Physical Review Letters 100(23):236405, 2008. https://doi.org/10.1103/PhysRevLett.100.236405. [24] T. M. G. Mohiuddin, A. Lombardo, R. R. Nair, et al. Uniaxial strain in graphene by Raman spectroscopy: G peak splitting, Grüneisen parameters, and sample orientation. Physical Review B 79(20):205433, 2009. https://doi.org/10.1103/PhysRevB.79.205433. [25] F. Ding, H. Ji, Y. Chen, et al. Stretchable graphene: a close look at fundamental parameters through biaxial straining. Nano Letters 10(9):3453–3458, 2010. https://doi.org/10.1021/nl101533x. [26] A. Contreras-Astorga, V. Jakubskỳ, A. Raya. On the propagation of Dirac fermions in graphene with strain-induced inhomogeneous Fermi velocity. Journal of Physics: Condensed Matter 32(29):295301, 2020. https://doi.org/10.1088/1361-648X/ab7e5b. [27] D. J. Fernández C. SUSY quantum mechanics. International Journal of Modern Physics A 12(01):171–176, 1997. https://doi.org/10.1142/S0217751X97000232. [28] D. J. Fernández C., N. Fernández-García. Higher-order supersymmetric quantum mechanics. AIP Conference Proceedings 744(1):236–273, 2004. https://doi.org/10.1063/1.1853203. [29] L. M. Nieto, A. A. Pecheritsin, B. F. Samsonov. Intertwining technique for the one-dimensional stationary Dirac equation. Annals of Physics 305(2):151–189, 2003. https://doi.org/10.1016/S0003-4916(03)00071-X. 29 https://doi.org/10.1088/1751-8113/47/28/285302 https://doi.org/10.1063/1.4898184 https://doi.org/10.1088/1751-8121/ab3f40 https://doi.org/10.1103/PhysRevB.102.115429 https://doi.org/10.1088/1361-6633/aa74ef https://doi.org/10.1002/pssr.201800237 https://doi.org/10.1038/nnano.2013.161 https://doi.org/10.1038/nmat3783 https://doi.org/10.1103/PhysRevA.85.013818 https://doi.org/10.1103/PhysRevLett.119.113901 https://doi.org/10.1103/PhysRevA.94.063831 https://doi.org/10.1103/PhysRevA.96.013813 https://doi.org/10.1103/PhysRevA.84.021806 https://doi.org/10.1103/PhysRevLett.110.013903 https://doi.org/10.1038/nphoton.2012.302 https://doi.org/10.1103/PhysRevLett.110.266801 https://doi.org/10.1088/1367-2630/11/9/095003 https://doi.org/10.1103/PhysRevLett.100.236405 https://doi.org/10.1103/PhysRevB.79.205433 https://doi.org/10.1021/nl101533x https://doi.org/10.1088/1361-648X/ab7e5b https://doi.org/10.1142/S0217751X97000232 https://doi.org/10.1063/1.1853203 https://doi.org/10.1016/S0003-4916(03)00071-X Acta Polytechnica 62(1):23–29, 2022 1 Introduction 2 Strain in photonic graphene 2.1 Tight-binding model 2.2 Uniform strain 2.3 Non-uniform strain 3 Supersymmetric quantum mechanics: matrix approach 4 Photonic graphene under strain and position-dependent gain and loss 4.1 Photonic Graphene with a single mode 4.2 Photonic Graphene with two modes 5 Summary Acknowledgements References