1. Introduction
In 1926, Schrödinger published his papers on wave mechanics, introducing the equation named after him [
1]. The time-dependent version [
2] has the form
(our discussion will be restricted to one dimension and systems with analytic solutions, i.e., potentials
V that are at most quadratic in the position variable), and
is a complex function of space and time. Shortly after this, Madelung [
3] found a reformulation in terms of two real equations that have formal similarity with equations known from hydrodynamics. For this purpose, he used the polar form
for the complex wave function, where the amplitude is expressed in terms of the probability density
and the phase essentially depends on the function
that has the dimension of action. The two equations are the continuity equation
and a modified Hamilton–Jacobi equation
with the so-called quantum potential
Equations (
3)–(
5) are also the formal basis of Bohmian mechanics [
4,
5,
6]; however, Bohm’s interpretation is ontologically quite different. While Madelung stuck to the conventional interpretation of quantum mechanics, Bohm introduced the concept of “hidden” variables that should convert quantum mechanics to a deterministic theory. The deterministic aspect is represented by the so-called Bohmian trajectories that are obtained via integration of the velocity field
that appears in the continuity Equation (
3). These trajectories are assumed to be real geometric paths of quantum particles that actually exist. However, no experiment was able to detect these trajectories. To the contrary, the experiments of last year’s Nobel laureates, those of Aspect [
7,
8,
9] in particular, contradicted the existence of such physical trajectories. Moreover, we could show theoretically [
10] that there is an ambiguity in the derivation of these trajectories that can be eliminated; but this inevitably leads to an interpretation of these trajectories in terms of descriptive statistics, thereby providing a complementary (but still probabilistic) aspect to the conventional interpretation of quantum mechanics in terms of statistical inference. Therefore, the formalism, particularly the computational aspect of Bohmian mechanics, is still very useful and has many fields of application, as shown, for example, in the interesting review article by Benseny et al. [
11].
As can be seen in Equation (
3), the time-evolution of the amplitude
of the wave function
contains its phase,
, and Equation (
4) for the phase contains the amplitude via
, so both equations are coupled.
In a recent paper [
12], we have shown how this coupling affects the uncertainties of position and momentum and their correlation, and particularly, how the momentum uncertainty can be split into contributions from the amplitude and from the phase. In this paper, we show the influences of a dissipative environment on these properties.
Let us first specify what we mean by environment and how this environment can be taken into account in our case. In physics, one usually does not consider the whole universe, only part of it that one calls the “system.” Ideally, we assume this system to be isolated from the rest of the universe. This is the case, e.g., in classical and in quantum mechanics. However, it can also be of interest to include interactions of this system with the rest of the universe, or at least parts of it. This rest is usually called the environment, the reservoir, a heat bath or something similar. These systems, which are no longer considered isolated, are therefore also called open systems, and there are different approaches to treating such systems.
The viewpoint of there being no artificial distinction between system and environment, but of considering both together as a closed, isolated system, is closest taken by the so-called system-plus-reservoir approach. In this case, the system of interest is explicitly coupled to a (large) number of environmental degrees of freedom (often represented by harmonic oscillators). In the end, however, the environmental degrees of freedom are averaged out, and only their influence on the system is considered. This type of approach is often called the Caldeira–Leggett model [
13,
14] but has also been considered independently by other authors (for more details, see, e.g., [
15,
16]). The more environmental degrees of freedom are taken into account (in the limit infinitely many, corresponding to the so-called Markovian assumption, which actually artificially breaks the time symmetry of the system’s evolution), the better the results.
As the system-plus-reservoir approach requires extensive calculations for a large number of external degrees of freedom, possibly exceeding even the potential of powerful computers when nonlinearities are involved, and the environmental degrees of freedom are averaged out in the end anyway, a different type of approach, the effective one, takes into account only the effect of the environment on the system of interest, but not the individual interactions with the environmental degrees of freedom.
There are various approaches in this direction in the literature [
17,
18,
19,
20,
21,
22] using modifications of the time-dependent Schrödinger Equation (
1), but most of them suffer serious shortcomings concerning their physical consequences with respect to irreversibility and correct dissipation of the system. We chose an approach that does not have these problems and start by breaking the time-reversal symmetry of the continuity Equation (
3) by adding a diffusion term, thereby changing it to a Fokker–Planck-type Equation (in position space called Smoluchowski equation). With an additional separation condition, this real equation can be split into two (conjugate) complex Schrödinger-type equations with complex logarithmic nonlinearity.
The nonlinearity vanishes on average, so no imaginary contribution to the energy appears, and normalization of the corresponding wave function is still guaranteed. However, the nonlinearity has a significant influence on the analytic behavior of the time-dependence of the exact Gaussian-shaped wave-packet solutions of this nonlinear equation. Therefore, also the time-dependence of the phase and amplitude will be influenced by the additional nonlinear term.
As in our previous paper without dissipative environment [
12], we again consider the uncertainties of position, momentum and their correlation; and the contributions of amplitude and phase of the wave function, but now analyze the effects of the environmental term on them.
We use these uncertainties to formulate their contributions to the energy of the system and the Heisenberg uncertainty product and their time-dependence. Using the parameters that connect the quantum system with a Brownian-motion environment, relations can also be established with thermodynamic properties of this environment.
In
Section 2, our complex hydrodynamic version of Madelung’s approach, proposed in [
23], will be introduced and applied in connection with the quantum uncertainties. In particular, the explicit form for generalized coherent states, i.e., the analytical Gaussian wave-packet solutions with time-dependent width will be shown.
The discussion in
Section 3 focuses on the interaction with an environment. For this purpose, a time-symmetry-breaking diffusion term is used in the continuity equation, and using a particular separation condition, a nonlinear modification of the time-dependent Schrödinger equation is obtained. The properties of the complex nonlinear additional term are also pointed out.
The consequences are shown for the corresponding modified Hamilton–Jacobi equation and for the equations of motion for the wave packet maximum and width. This also elucidates the influences of real and imaginary parts of the nonlinear term on the uncertainties of position and momentum and the contributions of amplitude and phase to these quantities. This is also investigated for the Gaussian wave-packet solutions that still exist despite the logarithmic nonlinearity.
These results provide the foundation for
Section 4. Based on them, the uncertainties are used to construct their contributions to the energy of the system and to the Heisenberg uncertainty product. In particular, the harmonic oscillator and the free motion, without and with environment, are compared for selected initial conditions, followed by a discussion about connections with thermodynamic properties of the environment, by way of parameters characterizing this environment.
Finally, in
Section 5 the results are summarized and conclusions drawn.
3. Open Systems with Irreversibility and Dissipation
There are different approaches to treating open systems classically and quantum mechanically (for a survey, see, e.g., [
16]). One possibility is to couple the system of interest to a bath of environmental degrees of freedom, e.g., harmonic oscillators where the number of these degrees of freedom must be very large in order to obtain physically sound results, and in the end, they are averaged out. This leads to extensive and costly calculations that can even go beyond the capacities of current computers when nonlinearities are involved.
A different viewpoint is taken by so-called effective approaches where the individual degrees of freedom of the environment are not taken into account explicitly, only their effect on the system of interest. Amongst those are attempts to modify the classical Lagrange–Hamilton formalism in a way that additional dissipative friction forces, usually linear proportional to velocity or momentum, can be incorporated into the equation of motion of the system. In general, this involves non-canonical transformations with corresponding challenges [
16,
17,
18]. A quantum mechanical version is then obtained via canonical quantization, leading to modified, often explicitly time-dependent but usually linear, modifications of the Schrödinger equation.
A different class of effective approaches circumvents the problems with non-canonical aspects and starts already on the quantum mechanical level by adding terms to the Hamiltonian operator that yield the above-mentioned friction forces, leading to nonlinear Schrödinger equations. One way of obtaining the additional terms is to require that these provide, according to Ehrenfest, the desired friction force in the equation of motion. In our Madelung picture, the equation that leads to the force is the modified Hamilton–Jacobi Equation (
3). Taking the gradient of this equation and using
provides a kind of Euler equation with substantial time-derivative
in a co-moving frame:
A friction force should appear as an additional term proportional to
(with negative sign) on the rhs of Equation (
38). As this requirement leaves room for ambiguity, there are several approaches in the literature to achieve this goal [
19,
20,
21,
22], but most of them suffer serious shortcomings and produce unphysical results.
Another problematic aspect is the fact that (
38) is related only to the real part of the Schrödinger equation, so the additional friction force in (
38) should originate from an additional real term in the modified Schrödinger equation. However, since real terms do not affect the Madelung equation representing the imaginary part of the Schrödinger equation, this would still remain a reversible continuity equation, contradicting the observation that dynamics of open system is irreversible, particularly if friction is involved.
Therefore, we use a different approach, starting in the Madelung picture with the continuity Equation (
2) and breaking the time-reversal symmetry by introducing an additional diffusion term:
This turns the reversible continuity equation into an irreversible Fokker–Planck-type equation in position space called the Smoluchowski equation, which is known from the description of Brownian motion. An equivalent description of Brownian motion in the trajectory picture via the Langevin equation involves the above-mentioned friction force that will also be compatible with our approach, as will be shown below.
The convection velocity field
of (
2) is now replaced by the total velocity field
with the additional diffusion velocity
, where
D is the diffusion coefficient characterizing the environment.
In conventional quantum mechanics, the continuity equation for
is obtained by combining the Schrödinger equation for
with its complex conjugate for
. Thus, the question arises of whether this procedure can also be reversed to obtain the complex modified Schrödinger equation corresponding to (
39) by separation of this equation. A way to achieve this without the diffusion term was shown by Madelung [
25] and Mrowka [
26]. Unfortunately, due to the diffusion term,
and
are coupled, and therefore, the above-mentioned method cannot be applied in general. However, there might be special conditions that can be fulfilled by the diffusion term to still allow separation.
One such separation condition is given by
that is also fulfilled particularly for Gaussian functions—those we consider in our analysis. The subtraction of the mean value of
is necessary to guarantee normalizability of the solutions of the corresponding Schrödinger equation, as the diffusion term leads to an imaginary contribution in the Hamiltonian, thereby turning it into a non-Hermitian one.
Separation of (
39), using (
41), leads to a nonlinear Schrödinger equation with complex logarithmic nonlinearity:
with
Due to the appearance of the imaginary part
, the nonlinear Hamiltonian is non-Hermitian. This usually leads to problems with the normalizability of the corresponding wave functions. In a similar approach to describe open dissipative systems, Gisin [
27,
28], via a Master equation for the wave function, also arrived at an imaginary term in the Hamiltonian, but enforced normalizability via subtraction of its mean value. In our case, this mean value consistently originates from the separation condition (
41). Gisin’s approach contains some ambiguities; and also his expression for the decay of the system’s energy disagrees with the one known, but it was a step in the right direction.
A nonlinear Schrödinger equation, containing only our imaginary contribution without the real (dissipative) term, was obtained independently by Beretta [
29,
30,
31] (but in the context of density operators) with the aim of describing non-equilibrium systems (without dissipation).
In an attempt to reduce the treatment of the quantum mechanical many-body wave function to a discussion in terms of a single-particle wave function, Oriols [
32] also arrived at an imaginary term in the corresponding Schrödinger equation. However, as admitted in [
11] concerning this imaginary term, the “numerical values are in principle unknown and need some educated guesses” [
32,
33].
Further details concerning the derivation of Equation (
42), its properties and connections with similar nonlinear approaches can be found in [
16].
As a result of the subtraction of the mean value of
, the mean value of the nonlinear term vanishes; therefore, the mean value of energy is still the mean value of the operators of kinetic and potential energies. However, the mean values are now calculated with the solution
of the nonlinear Schrödinger equation (Equation (
42)). The nonlinear term obviously influences the dynamics of phase and amplitude, and thus of the maximum and width of the wave packet. The details will be obvious when the equations of motion of these quantities are examined.
The additional nonlinear term also has a real contribution that is not arbitrary but fixed by the separation condition and has a clear physical meaning. A first idea of this meaning can be obtained by looking at the second of Madelung’s equations, the modified Hamilton–Jacobi equation. Due to the real part of the nonlinear term, it turns into
Taking the gradient of this equation now yields the Euler Equation (
38) with an additional friction term that is linearly proportional to momentum, as known from the Langevin equation, but without putting it in by the assumption of a corresponding “friction potential”, like in similar approaches [
19,
20,
21,
22] mentioned above:
Considering now the Gaussian solutions of the nonlinear Schrödinger Equation (
42) in the form (
21), the additional complex nonlinear term
W can be expressed as
where the negative gradient (in one dimension) yields
showing that its mean value is real and provides the friction force in the Ehrenfest equation; i.e.,
After inserting the Gaussian ansatz (
21) into the nonlinear Schrödinger Equation (
42), the corresponding equations of motion for maximum and width of the wave packet can be obtained as
The friction force in (
49) was already mentioned. Important for the uncertainties, however, is the Riccati Equation (
50) with the additional linear term
. Splitting (
50) into real and imaginary parts from the imaginary part follows:
or
where the definition
has been unchanged, as in the case without environment. This shows that
, which is crucial for the contribution of the
phase to the uncertainties, is changed by
, which originates from the imaginary part of
W, and thus from the Smoluchowski equation for the
amplitude!
On the other hand, the real part of (
50) leads to
which can be written with
and
expressed in terms of
as a modified Ermakov equation:
that is identical to (
27); only
is replaced by the correct reduced frequency
of the damped harmonic oscillator.
Here, it is exactly the other way round; the equation for , and thus the wave packet width that characterizes the amplitude, is influenced by , which originates from the real part of W, the one that corresponds to the modified Hamilton–Jacobi equation for the phase.
Thus, there is an additional coupling of phase and amplitude due to interaction with the environment.
5. Conclusions
Using a polar ansatz for the wave function
in terms of amplitude
and phase
, Madelung [
3] proposed a formulation in terms of two real equations that are formally similar to classical hydrodynamic equations, a continuity equation for the amplitude and a modified Hamilton–Jacobi equation for the phase, instead of the complex Schrödinger equation. These two equations are coupled, as the gradient of the phase occurs as a velocity (actually momentum) field in the continuity equation, and the amplitude occurs in the form of the so-called quantum potential
in the modified Hamilton–Jacobi equation.
In a recent paper, we have shown how the Madelung picture can be formulated in terms of complex quantities, not only allowing for the formulation in momentum space and beyond [
23], but also for its extension to other hydrodynamic properties [
24]. In position space (to which the discussion in this paper is restricted), the momentum operator and its corresponding complex hydrodynamic quantity,
, are of interest. The real part
only depends on the phase
; the imaginary part
only on the amplitude
. The situation changes when nonlinear functions of
are considered, such as for the kinetic energy, where the square of the momentum appears. The real part of the complex quantity corresponding to the kinetic energy does not only contain contributions from
, but also from
and its derivative. Although the mean value of
, like all mean values of the complex Madelung quantities, vanishes, this is not the case for the contributions of
to the real part of the kinetic energy.
Consequently, the uncertainty product of and has two contributions, one from the amplitude and one from the phase. The contribution from the amplitude provides the minimum uncertainty . If the uncertainty product has a larger value, this contribution originates from the phase and can be expressed in terms of the correlation of position and momentum uncertainties via their anti-commutator.
To illustrate this situation, we use Gaussian wave packets that are exact analytic solutions of the time-dependent Schrödinger equation for the systems we consider, and that are completely determined by their maximum and width and the corresponding dynamics. The time-evolution of the maximum is determined by the classical equation of motion for the given potential; the time-evolution of the width by a complex nonlinear Riccati equation for . The imaginary part of this equation determines the real part of that enters the phase of the wave function, and thus and . The imaginary part of can be expressed in terms of a new variable that is directly proportional to the wave packet width, . By inserting and in terms of and into the real part of the Riccati equation, the time-evolution of the wave packet width can be directly obtained from a real nonlinear equation for , the so-called Ermakov equation.
Next, the influence of a dissipative environment on these systems was considered. Amongst several possible approaches to treat such open systems, we chose an effective one that is compatible with our hydrodynamic Madelung picture. Since the interaction with a dissipative environment breaks the symmetry under time-reversal of the dynamics, this should also be reflected in the corresponding equations of motion. Unlike other approaches that are based on phenomenological friction forces, like the one in the Langevin equation of Brownian motion, something that could be related to the modified Hamilton–Jacobi equation (Equation (
4)), our approach starts from the continuity Equation (
3), breaking the time symmetry by introducing an irreversible diffusion term, leading to a Fokker–Planck-type equation, which in position space is a Smoluchowski equation. The attempt to separate this equation for
into two equations depending only on
or
is not successful in general due to the coupling of
and
via the diffusion term. However, introducing an additional separation condition, where
is related to
, permits separation. Although this separation condition restricts the possible solutions of the resulting modified Schrödinger equation, it still enables Gaussian wave packets, the ones we consider in this paper, and also enables normalization of the solutions, though the diffusion term leads to a (nonlinear) imaginary term in the Hamiltonian operator, thereby changing it to a non-Hermitian one.
Due to separation, an additional real contribution is introduced into the Hamiltonian that, according to Ehrenfest, provides the desired friction force proportional to velocity or momentum, known from the Langevin equation, but, as a consequence of the separation condition, it is not introduced as a starting requirement for the construction of the additional interaction term in the Hamiltonian. Consequently, also in the second Madelung equation, the modified Hamilton–Jacobi equation, this friction force proportional to momentum appears.
As the mean value of the additional nonlinear term vanishes,
, no additional observable term appears in the mean value of the Hamiltonian, and thus in the energy, i.e.,
. However, these mean values are now evaluated with the solutions of the nonlinear Schrödinger equation (Equation (
42)), leading to different time-dependence of the mean values of the energy and the uncertainties.
That also means that
and
, as functions of
,
and
, or
and
, respectively, are unchanged, but the time dependence of
and
that determines that of
,
and
, definitely depends on the additional terms from
. The imaginary part
, originating from the Smoluchowski equation (Equation (
39)) for
, enters the imaginary part of the Riccati equation (Equation (
51)), changing
according to (
52). As
is the quantity that enters
, the terms that are responsible for the addition to the minimum uncertainty product via
are changed too.
The real part of the additional term, , corresponding to the modified Hamilton–Jacobi equation for the phase, adds a linear term to the real part of the Riccati equation, thereby changing the Ermakov equation for the width of the wave packet and the time-dependence of the amplitude.
As examples, the harmonic oscillator and the free motion, without and with environmental interaction, have been considered. For the harmonic oscillator, the interaction with the environment provides an additional contribution to the energy that originates from the quantum uncertainties, causing it to be larger than without environment. This is similar to the situation of a Brownian particle in a bath of temperature T, where the final energy is due to the back-transfer of energy from the bath via stochastic, i.e., statistic, force, the average value of which vanishes, similarly to , the mean value of the imaginary part of our complex momentum. The mean value of does not vanish, similarly to the average value of the square of the stochastic force that supplies the final energy of the Brownian particle.
For the free motion, the interaction with the environment causes a bifurcation, so there are two states with different initial ground state energies and different uncertainty products.
The parameter that connects the quantum system via the Smoluchowski equation (Equation (
39)) with the environment is the diffusion coefficient
D. For different choices of
D, different environments can be taken into account. By choosing a Brownian motion environment and taking into account for this case the Einstein relation
, the excess energy of the damped harmonic oscillator and the difference between the lowest energies of the two states of the damped free motion, originating from bifurcation due to broken time-reversal symmetry, can be expressed in terms of the parameters of the environment, particularly in terms of its temperature
T.
Taking the time-dependent Schrödinger equation for the free motion as a diffusion equation with imaginary diffusion coefficient allows formal similarity between the imaginary part of the additional nonlinear term and the definition of entropy in statistical mechanics; however, in the quantum mechanical context, the imaginary unit appears.
In the paper concerning the uncertainties in our complex Madelung picture [
12], we considered position and momentum space. Treating the corresponding open systems with our nonlinearity (
43), the gradient of this term is proportional to momentum. In momentum space, however, the gradient of the logarithm of the wave function with respect to momentum would provide a friction force proportional to position, i.e., a totally different physical situation. Therefore, the treatment of open systems in momentum space needs a formally different approach that is beyond the scope of this paper but will be considered in forthcoming work.