1 Introduction
Graphene is a single layer of carbon atoms arranged in a hexagonal lattice pattern, often described as a ”honeycomb” structure. The 2-D material graphene was first successfully isolated by Novoselov and Geim et al. [33].
In 2010, Geim and Novoselov were awarded Nobel Prizes in Physics for their groundbreaking
experiments regarding graphene.
Due to graphene’s unique electrical, electromagnetic,
and optical characteristics, it has attracted widespread attention, leading to the design of
many new systems and equipment with graphene.
For example, graphene has
played a prominent role in the design of organic light-emitting
diodes, solar cells, antennas, and invisibility cloaks (cf. [12, 5, 40]).
In the past two decades, in addition to many great achievements on the physical research on graphene and graphene-based devices, another hot topic has been the numerical simulation of graphene.
The famous Kubo formula [13] gives the expression of graphene conductivity, which is a function
of many physical parameters such as wavelength, chemical potential, and temperature.
With the development of
computational electromagnetic in the past, such as the finite difference time-domain (FDTD) method (e.g., [14, 16, 19, 27, 39]) and
the finite element method (FEM) (e.g., papers [4, 7, 8, 9, 23, 35, 17, 37, 42], and books [11, 30, 24]), many robust and efficient numerical methods
have been proposed to simulate the electromagnetic response of graphene related devices.
The numerical approaches for modeling graphene can be classified into two big categories:
(1) Treating graphene as a thin plate with finite thickness [34], and
converting its surface conductivity into a volumetric conductivity;
(2) Taking graphene as a zero-thickness sheet [31].
Due to the
easy realization of (1), many published papers and commercial software model graphene
as a thin sheet with a finite thickness [6, 43]. However, direct discretization of graphene
with a small finite thickness results in extremely fine grids around graphene. This leads to extremely small time steps
for time-domain simulations with explicit schemes, which
consume enormous amount of memory storage and CPU time.
The interesting physical research on graphene and graphene-based
devices has inspired mathematicians to conduct mathematical analysis [1, 15, 21] and modeling of graphene
(e.g., [20, 29, 38, 44] and references therein).
Here we are interested in simulating the surface plasmon polaritons (SPPs) generated on the graphene surface,
since SPPs provide a window into the nano-optical, electrodynamic response of their host material and its dielectric environment [41].
The plasmons occur in the highly sought after terahertz to mid-infrared regime. Note that
terahertz waves are used in a variety of applications such as nondestructive analysis, since terahertz waves can be used to analyze the internal structure of objects without damaging them. For example, THz cameras can be used to see what is inside sealed packages. Also, terahertz waves can be found in military applications, e.g.,
terahertz sensors are used in military communication, detecting biological and explosive warfare agents, and inspecting concealed weapons. However, because of practical difficulties in exciting and detecting the SPP waves in
graphene, numerical simulation of wave interactions with graphene materials plays a very important role in designing functional components with graphene.
Considering the disadvantage of FDTD methods in handling the complex geometry, which happens quite often in graphene devices,
we recently proposed and analyzed some finite element time-domain (FETD) methods for graphene simulation [45, 22, 18, 26].
In [45, 22], we treated the graphene with some thickness (though very thin); while in [26], we successfully developed a new FETD method for simulating surface plasmon polaritons (SPPs) on graphene sheets with zero-thickness.
We like to remark that the works of [29, 38] are based on frequency-domain finite element methods.
The work of Wilson, Santosa, and Martin [44] is for
2-D time-domain SPP models written as an integro-differential equation in time. In [32], Nicholls et al. proposed the so-called
high–order perturbation of surfaces algorithms for simulating the plasmonic response of a perfectly flat sheet of graphene. They solved the Helmholtz equations by using the dispersive Drude model for the
surface current.
In this paper, we propose a novel FETD method by treating the graphene sheet as zero-thickness.
The new method is based on a reformulated system of governing equations for graphene with electric and magnetic fields as unknowns.
Compared to our previous work [26], the new system does not contain the induced current explicitly. Hence, our new method is more efficient in memory and computational cost.
The rest of the paper is organized as follows. In Section 2, we first
present and reformulate the time-domain governing equations for modeling the surface plasmon polaritons
on graphene sheets. Then we prove the stability for the model. In Section 3, we propose a leapfrog
time stepping finite element scheme for solving the graphene model.
In Section 4, we present extensive numerical results to demonstrate the effectiveness of our method for simulating the propagation of surface plasmon polaritons on various complex graphene sheets.
We conclude the paper in Section 5.
2 The governing equations and stability analysis
In our previous papers [45, 22],
we treated the graphene as a homogenized material with an effective permittivity and a small thickness.
By ignoring the interband conductivity, we have the model governing equations for simulating surface plasmon propagation on graphene [45, (2.7)-(2.12)]:
|
|
|
(2.1) |
|
|
|
(2.2) |
|
|
|
(2.3) |
where we denote the electric field , magnetic field ,
for an imposed magnetic source function, (as denoted in [45])
for the induced intraband surface current in graphene,
and for the vacuum permittivity and permeability,
the positive constant for the relaxation time,
and the positive constant for the graphene surface conductivity.
Moreover, we assume that the physical domain is a bounded Lipschitz polygonal domain in with boundary , and represents the graphene sheet embedded in .
Also we define the 2-D curl operators as
and .
To complete the model,
we assume that (2.1)-(2.3) are subject to the simple perfectly conducting (PEC) boundary condition:
|
|
|
(2.4) |
and the initial conditions
|
|
|
(2.5) |
where denotes the unit outward normal vector on ,
and are some properly given functions.
Recently, we [26] successfully simulated surface plasmon propagation on graphene by treating it
as zero-thickness sheet (i.e., appearing as a curve in our 2-D simulations, e.g., Figures 1 and 3 shown later),
with the following boundary conditions given on the graphene interface [3, Fig.1]:
|
|
|
(2.6) |
|
|
|
(2.7) |
where and represent the magnetic fields above and below the interface, respectively,
denotes the unit normal vector of the interface pointing upward, and and
are the unit outward normal vectors from the top and bottom
of the interface. Finally, we adopt the 2-D cross product notation
. We like to remark that (2.6)-(2.7) imply that the tangential electric field is continuous across the interface,
and the jump of the tangential component of the magnetic field across the interface is equal to the surface current.
In [26], using integration by parts and the interface conditions (2.6)-(2.7),
we derived the following weak formulation for simulating graphene as an interface:
Find the solution
|
|
|
such that (cf. [26, (2.9)-(2.11)]):
|
|
|
(2.8) |
|
|
|
(2.9) |
|
|
|
(2.10) |
hold true for any test functions ,
and . To obtain (2.8), we use integration by parts over and the boundary condition (2.7).
Here we denote for the inner product over
, and for the inner product on .
Finally, we adopt the standard Sobolev space notation
|
|
|
To derive the governing equations for our new numerical method, we differentiate (2.8) with respect to and then multiplying the result by . This gives
|
|
|
(2.11) |
Adding (2.11) and (2.8) together, and using (2.10) with , we have
|
|
|
(2.12) |
To develop a more efficient numerical method later, we replace in (2.12) by (2.9) with and obtain the following new weak formulation: Find the solution
,
,
such that
|
|
|
|
|
|
(2.13) |
|
|
|
(2.14) |
hold true for any test functions and .
The initial conditions for the problem (2.13)-(2.14) are as follows:
|
|
|
(2.15) |
To simplify the notation, we denote the norm of in
as , and the norm of on
as .
Theorem 2.1.
Denote the energy
|
|
|
|
|
(2.16) |
|
|
|
|
|
Then we have the following continuous stability:
|
|
|
(2.17) |
where the constant depends only on the parameter .
Proof. To make our proof easy to follow, we split it into several major parts.
(I) Choosing in
(2.13), we have
|
|
|
|
|
|
(2.18) |
Taking the time derivative of (2.2), we obtain
|
|
|
(2.19) |
Replacing in (2.18) by (2.19), we have
|
|
|
|
|
|
(2.20) |
(II) By choosing in (2.14), then
replacing by (2.19), we have
|
|
|
|
|
(2.21) |
|
|
|
|
|
|
|
|
|
|
Adding (2.20) and (2.21) together, and using (2.2), we obtain
|
|
|
|
|
(2.22) |
|
|
|
|
|
|
|
|
|
|
which can be rewritten as
|
|
|
|
|
(2.23) |
|
|
|
|
|
(III) To bound the terms and on the right-hand side (RHS) of (2.23),
we first choose in
(2.8)-(2.10), respectively. Adding the results together, we obtain
|
|
|
(2.24) |
Multiplying (2.24) by , we have
|
|
|
(2.25) |
Similarly, taking the time derivative of (2.8)-(2.10), then choosing respectively, and
adding the results together, we obtain
|
|
|
(2.26) |
Multiplying (2.26) by , we have
|
|
|
(2.27) |
Now adding (2.23), (2.25), and (2.27) together, we have
|
|
|
|
|
(2.28) |
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
(IV) Using the Young’s inequality, we can bound the last three RHS terms of (2.28) as follows:
|
|
|
(2.29) |
|
|
|
(2.30) |
|
|
|
(2.31) |
Substituting the above estimates (2.29)-(2.31) into (2.28), and dropping the last three non-negative terms on the left hand side of (2.28), we obtain
|
|
|
|
|
(2.32) |
|
|
|
|
|
|
|
|
|
|
The proof is complete by applying the Gronwall inequality to (2.32).
3 The leapfrog finite element scheme and its analysis
To design a finite element method, we partition the physical domain
with as an internal boundary
by a shape regular triangular mesh with maximum
mesh size . Without loss of generality, we
consider the following
Raviart-Thomas-Nédélec (RTN)’s mixed spaces and
on triangular elements [24, 30]: For any ,
|
|
|
|
|
|
where denotes the space of polynomials
of degree less than or equal to , and represents the space of homogeneous polynomials of degree .
To handle the PEC boundary condition (2.4), we introduce the subspace
|
|
|
To construct the fully discrete finite element scheme, we partition the time interval uniformly by points , where denotes the time step size.
Now we can construct the following leapfrog time stepping scheme:
Given proper initial approximations of , , for any , find , such that
|
|
|
|
|
|
(3.1) |
|
|
|
(3.2) |
hold true for any test functions .
Here we adopt the following central difference operators and averaging operator in time: For any time sequence function ,
|
|
|
Corresponding to the finite element spaces and ,
we denote
and for the standard Nédélec interpolation
in space and the standard projection onto
space , respectively.
Furthermore, the following interpolation and projection errors hold true (cf. [24, 30]):
|
|
|
(3.3) |
|
|
|
(3.4) |
where denotes the norm for the Sobolev space , and
is the norm for the Sobolev space
|
|
|
The initial conditions (2.15) are discretized as follows:
|
|
|
(3.5) |
|
|
|
(3.6) |
|
|
|
(3.7) |
where we used the Taylor expansion and the governing equation (2.2).
The implementation of the scheme (3.1)-(3.2) is quite simple. At each time step, we first solve (3.2) for ; then solve (3.1) for .
Of course, at the first time step when , we need to use the initial conditions (3.5)-(3.7).
It can be seen that the coefficient matrix at each time step is symmetric and positive definite. Hence, the solvability of our scheme (3.1)-(3.2) is guaranteed.
3.1 The stability analysis
To prove the discrete stability for our scheme (3.1)-(3.2),
we denote for the wave propagation speed,
and recall the standard inverse estimate:
|
|
|
(3.8) |
and the trace inequality:
|
|
|
(3.9) |
where the positive constants and are independent of the mesh size .
Theorem 3.1.
Denote the discrete energy for the scheme (3.1)-(3.2):
|
|
|
|
|
(3.10) |
|
|
|
|
|
Then under the time step constraint:
|
|
|
(3.11) |
we have the following stability: For any ,
|
|
|
(3.12) |
where the positive constant is independent of and .
Proof. To make our proof easy to follow, we divite it into several major steps.
(I) Choosing in (3.1), and using the following identities
|
|
|
(3.13) |
|
|
|
(3.14) |
with , we have
|
|
|
|
|
|
|
|
|
|
|
|
(3.15) |
Taking
in (3.2), and using (3.14) with , we obtain
|
|
|
(3.16) |
|
|
|
Taking
in (3.2), we obtain
|
|
|
(3.17) |
Adding (3.15), (3.16), and (3.17) together, then summing up the result from to
any , we have
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
(3.18) |
(II) By the definition of the discrete energy and dropping the non-negative term
on the left hand side of (3.18), we can rewrite (3.18) as follows
|
|
|
|
|
|
|
|
|
|
|
|
(3.19) |
Now we just need to estimate each in (3.19). Using the inverse estimate (3.8), it is easy to see that
|
|
|
(3.20) |
and
|
|
|
(3.21) |
Using the trace inequality (3.8), we have
|
|
|
(3.22) |
Similarly by the inverse estimate (3.8) and the Cauchy-Schwarz inequality, we have
|
|
|
|
|
(3.23) |
|
|
|
|
|
It is also easy to obtain the following estimates:
|
|
|
|
|
(3.24) |
|
|
|
|
|
and
|
|
|
(3.25) |
The estimates of and need a special treatment, since can not be controlled by . In the next major part, we will carry out the estimates of and .
(III) Note that
|
|
|
|
|
(3.26) |
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
Using (3.26) and the inequality , we have
|
|
|
|
|
(3.27) |
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
By the same technique, we have
|
|
|
|
|
(3.28) |
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
Substituting all the above estimates of into (3.19) and collecting like terms, we obtain
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
(3.29) |
(IV) Now all the right hand side terms of (3.29), except and , can be controlled by choosing small enough and using the discrete Gronwall inequality. To bound these two terms, squaring the following identity
|
|
|
|
|
(3.30) |
|
|
|
|
|
we have
|
|
|
|
|
(3.31) |
|
|
|
|
|
By the same argument, we have
|
|
|
(3.32) |
To bound , taking in (3.2), we obtain
|
|
|
|
|
(3.33) |
|
|
|
|
|
which leads to
|
|
|
(3.34) |
We can use similar techniques
to bound the terms and in (3.29), even though they are given source functions and we can keep them as they are. When is even, we have
|
|
|
(3.35) |
When is odd, we have
|
|
|
(3.36) |
Substituting the estimates of (3.31), (3.32), and (3.34)-(3.36) into (3.29), and
choosing small enough, e.g.,
|
|
|
|
|
|
(3.37) |
which is equivalent to (3.11), we complete the proof by the discrete Gronwall inequality.
3.2 The error estimate
To carry out the error estimate of our scheme (3.1)-(3.2), we split the solution error for as follows:
|
|
|
(3.38) |
where for simplicity we denote . Similarly, we can define the solution error
for as follows:
|
|
|
(3.39) |
Integrating (2.13) with respect to from to , then divide the result by , we obtain
|
|
|
|
|
|
|
|
|
|
Integrating (2.14) with respect to from to , then divie the result by , we have
|
|
|
(3.41) |
Now subtracting (3.1) from (3.2) with ,
(3.2) from (3.41) with ,
using the error notation we introduced, and after some lengthy but straightforward algebra, we can obtain the error equations:
|
|
|
|
|
(3.42) |
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
and
|
|
|
|
|
|
(3.43) |
Note that the error equation (3.42) has the first five terms in the same form as scheme (3.1), and the rest terms are extra error terms due to the time approximation, and projection or interpolation.
Furthermore, the error equation (3.43) has the first two terms in the same form as scheme (3.2), and the rest three terms are error terms due to the time approximation, and projection or interpolation.
We want to remark that those extra error terms are by the interpolation error estimate (3.3) and the projection error estimate (3.4). Following the stability proof (due to its technicality, we skip the proof details), we have the following error estimate for our scheme (3.1)-(3.2):
|
|
|
(3.44) |
where is the degree of our finite element spaces and .