1. Introduction
Over the past years, the study of classical and/or quantum systems that may be described by the same mathematical formalism has received considerable attention of physicists and continues to be a living and interesting research area of modern physics. This interest is because the elucidation of similarities in the behavior of systems studied by different areas of physics frequently allows the knowledge of each area to help the understanding of the other.
In recent years, the phenomenon of superconductivity has been the focus of considerable theoretical and experimental study in the literature. The origin of this great interest is due to its potential application in various branches of physics [1-6]. This phenomenon was first observed in Leiden by Onnes in 1911 [7] when he noted that the resistance of a rod of frozen mercury suddenly drops to zero when cooled to the boiling point of helium, 4.2 K, and in 1933 [8], Meissner and Ochsenfeld discovered that this phenomenon expels the magnetic field from within the superconductor. This discovery is nowadays known as the Meissner-Ochsenfeld effect. Later, other important contributions to the study of superconducting materials were given by some researchers [9-13]. However, a great deal of our knowledge on superconductivity can be obtained from the London equations [9,13].
With the development of nanometer techniques and microelectronics, classical and quantum effects of mesoscopic or nanoscale circuits have attracted a lot of interest from physicists [14-20]. In the study of mesoscopic circuits, an LC (inductance L and capacitance C) circuit represents a fundamental cell. To our knowledge, the quantization of this circuit was first performed by Louisell in the 1970s [21]. This author has studied the quantum effects of a nondissipative mesoscopic LC circuit with a source and expressed its fluctuations in the vacuum state. Based on Louisell’s work and with the progress of quantum information and quantum computation, many articles concerned with mesoscopic or nanoscale circuits have been published in the literature [1-5,22-25].
The connection of superconductors with a mesoscopic LC circuit can be made through the Josephson junctions. In fact, one can consider Josephson junctions connected to an inductance by a capacitor [1,2,6]. So, it is of high interest to academics to study mesoscopic LC circuits from both classical and quantum viewpoints. Yet, in recent years physicists have found that a practical solid quantum computer might be built by using some solid-state devices, for example, taking a Josephson junction as its core [1,2]. Here, it is worth remembering that a Josephson junction is composed of two superconductors weakly connected by a thin layer of insulating material [1,2,6].
In the present paper, stimulated by a connection between superconductors and mesoscopic LC circuits, we investigate the behavior of a London superconductor and a mesoscopic LC circuit with a time-varying inductance increasing exponentially and a time-dependent capacitance decreasing exponentially from the classical and quantum viewpoints. Notably, we find that the behavior of these systems is identical and can be described by the Caldirola-Kanai Hamiltonian. Furthermore, using the dynamical invariant method developed by Lewis and Riesenfeld [26] and Fock states, we easily solve the time-dependent Schrödinger associated with this Hamiltonian and write its solutions in terms of a special solution of the Milne-Pinney equation [27,28]. Finally, we use Fock states to calculate some quantum properties of these systems, such as the expectation values of the charge and magnetic flux, their quantum variances, and the corresponding Heisenberg uncertainty principle.
This paper is organized as follows. In Sec. 2, we discuss the classical equivalence of the London superconductor and of the time-dependent mesoscopic LC circuit. The quantum equivalence is presented in Sec. 3. In Sec. 4, we conclude the paper with a short summary.
2. Classical description
2.1. London Superconductor
In this subsection, we discuss the classical behavior of the London superconductor. In order to do so, let us first consider, for the sake of clarity on the exposition of the mathematical formalism, the classical Maxwell’s equations with free charge and currents sources which can be written in the form
where ρ is the total density charge. The constitutive equations relating the fields are given by
where
The electromagnetic equations of the superconductor, that is, the London’s equations, are given by [9,13,29]
where
is the London penetration depth. In the above expressions,
where
with σ being the electric conductivity. Then, differentiating Eq. (9) with respect to time and using the Maxwell and London equations as well as the continuity equation (5), we obtain the equation of motion for the total charge within a certain volume of the superconductor as
where the dots stand for time-derivatives,
where A and δ are constants to be determined by the initial conditions and Ω is given by
with Ω2 > 0 (oscillatory solutions). The Eq. (11) can be easily derived from the time-dependent classical Hamiltonian
where q and Φ are canonical variables with Φ being the magnetic flux. This Hamiltonian is the well-known Caldirola-Kanai Hamiltonian, which has been used in the literature to study time-dependent systems in various areas of physics [19,30-35]. It is also easily verified that for this case, the classical magnetic flux is given by
which can be rewritten as
where
represents the inductance of the London superconductor. What is more, Eq. (15) yields
which is the Faraday’s law for the London superconductor. Therefore, the above results give a complete classical description of the London superconductor.
2.2. Time-dependent mesoscopic LC Circuit
In the present subsection, we are interested in studying the classical behavior of the time-dependent mesoscopic LC circuit. The classical scheme of this circuit is well-known. In the present case, it consists of an inductance L(t) and a capacitance C(t). The time-dependent capacitance enters the total voltage as q
1
/C(t), and the time-dependent inductance induces a magnetic field with the flux Φ1 = L(t)i
1, which contributes to the voltage, that is, Faraday’s law. Here,
where
where η is a positive constant and L 0 = L(0) and C 0 = C(0). So, we can rewrite the equation (19) as
where
The equation of motion (21) can be derived from the Hamiltonian
where Φ1 is the magnetic flux which is given by
Here, we note that the Hamiltonian (22) is similar to the Hamiltonian (14). From Eq. (23) we get Faraday’s law as
Furthermore, the solution of Eq. (21) is given by
where the constants B and ξ are determined by the initial conditions and Ω2 1 > 0 is the modified frequency of the circuit, which is given by
At this stage, we let us observe that the mathematical formalism for describing the classical behavior of the London superconductor and the time-varying mesoscopic LC circuit with inductance and capacitance modulated exponentially at a constant rate is identical. In both cases, the equations of motion are governed by a standard damped harmonic oscillator [Eqs. (11) and (21] which can be obtained from similar Hamiltonians [Eqs. (14) and (22]. These systems also possess similar expressions for the magnetic flux [Eqs. (16) and (23], inductance [Eqs. (17) and (20] and Faraday’s law [Eqs. (18) and (24]. In order to push this analogy even further, we make the following correspondence: σ/ε ⇔ η and ε ⇔ L 0. However, the equivalence of these two systems is not complete. In fact, the dispersion relations ω and ω 0, which are inherent to each physical system, differ, and as a consequence, the behavior of the electromagnetic oscillations of each system is different. On the other hand, in the absence of dissipation, that is, σ = 0 and η = 0, the Hamiltonians (14) and (22) reduce to that of a time-independent harmonic oscillator with ε and L 0 playing the role of the mass of the standard time-independent mechanical oscillator.
3. Quantum description
In order to obtain the quantum description of our timedependent mesoscopic LC circuit or, equivalently, of the London superconductor, we must solve the time-dependent Schrödinger equation associated with the Hamiltonian (22) or Hamiltonian (14), respectively. Here, we consider the Hamiltonian (22). The time-dependent Schrödinger equation associated with this Hamiltonian is
where the charge q
1 and the magnetic flux Φ1 now are canonical operators satisfying the commutation relation
the solutions of the time-dependent Schrödinger equation (27) can be written in terms of orthonormalized eigenstates
and phase functions β n (t) as
where the λ n are time-independent eigenvalues and the phase functions β n (t) are derived from the equation
with the orthonormality condition
where ρ(t) is a time-dependent real function satisfying the Milne-Pinney equation [19,27,28]
with L(t) given by equation (20).
Next, our task is to solve the eigenvalue Eq. (29). In order to do this, we will use the Fock state base since, as it is wellknown, the quantum behavior of some quantum systems, in particular quantum harmonic oscillator-type systems, is more obvious in Fock states, which are states with specific numbers of energy quanta. Then, let us introduce annihilation and creation-type operators a(t) and a †(t) defined by [19,26,37]
with
In terms of these operators, the invariant (32) can be rewritten as
From the Eqs. (36) and (37) we see that the eigenvalue equation for I(t) (see Eq. (29)) can also be solved exactly, just as for the harmonic oscillator in the time-independent case by using the Fock states |n,ti. So, defining the Hermitian number operator by N = a
†
a so that
From Eq. (38) we see that the eigenstates of I(t) are also eigenstates of N and vice versa.
In what follows, we want to find the phase functions given by Eq. (31). By making the change |φ n ,ti → |n,ti and after performing some basic calculations, we get that
We now consider a particular solution of the Milne-Pinney Eq. (33) given by [19,32]
For this case, Eq. (42) reduces to
Therefore, we can write the solutions of the Schrödinger Eq. (27) as
with β
n
(t) given by Eq. (44). The general solution to the Schrödinger Eq. (27) can be written as
Next, we use the Fock states to calculate some quantum properties for the quantized mesoscopic LC circuit. To this end, we use the Eqs. (39), (40), and (41). After a little algebra, we find that
By using the above expressions, we find the quantum variances as
Multiplication of (50) by (51) yields
which represents the uncertainty principle for our system. By using the particular solution (43), the uncertainty principle (52) reduces to
From (53), we see that the uncertainty principle does not depend on time and that its value becomes larger when the dissipation, that is, η increases. What is more, the above results give us the quantization of the charge, the magnetic flux, and the uncertainty principle for our time-dependent mesoscopic LC circuit. Now it is worth noticing that in the absence of the dissipation, η = 0, the expression (53) becomes
which is similar to the uncertainty principle of a harmonic oscillator with frequency
Finally, we observe that for the London superconductor, the quantum description can be carried out following the same procedure used for the time-dependent nanoscale LC circuit case since, as we have demonstrated previously, the mathematical framework to study the behavior of both systems is identical.
4. Summary
In this work, we have presented a simple procedure to analyze the classical and quantum behavior of the London superconductor and of a mesoscopic LC circuit with an inductance and a capacitance varying exponentially with time at a constant rate. Notably, we have found that the behavior of these two systems is equivalent, both classically and quantum mechanically. We also have shown that this behavior can be mapped into, in both cases, a standard damped harmonic oscillator which is governed by the well-known Caldirola-Kanai Hamiltonian. Yet, we have noted that