Damped Motion in Classical and Quantum Mechanics - Comments

In summary: From a variational principle for which the Hamiltonian is the energy, you never get dissipative equations. And if you change the Lagrangian such that it produces the right equations, and you don't introduce explicit time dependence (which would be unnatural in a system where there are no external forces) the Hamiltonian will typically be zeros. Thus, from the point of view of the equations of motion, the system is in equilibrium.
  • #1
hilbert2
Science Advisor
Insights Author
Gold Member
1,598
605
hilbert2 submitted a new PF Insights post

Damped Motion in Classical and Quantum Mechanics
dampedforces.png


Continue reading the Original PF Insights Post.
 
  • Like
Likes Buffu, Mentz114 and Greg Bernhardt
Physics news on Phys.org
  • #2
I'm glad this subject is being discussed. The world would not function the way it does without friction, viscosity and various kinds of resistance.
In fact it is very difficult to eliminate these effects and any realistic modelling must include them.
But most of the physical modelling ignores dissipation and leads to strange and even ridiculous conclusions.

If a marble is set moving in a bowl it will keep moving forever without some kind of dissipative effect - in fact one could say that the experiment has no outcome unless there is friction. Dissipation creates definite outcomes and without including it in models one soon meets conundra like the 'measurement problem'.
 
  • #4
The way described in this Insight article is not the correct way to include dissipation into the description of physical systems; it works well only in a few toy examples. I'd like suggest that the author retracts the article and makes himself better informed before reposting an improved version.

For a system of ##n## oscillators with ##n##-dimensional position vector ##q##, the correct dissipative version of the conservative equation for nonlinear (Lagrangian or hamiltonian) oscillators dynamics ##M\ddot q+\nabla V(q)=0## is given by ##M\ddot q+C(q,\dot q)\dot q +\nabla V(q)=0##. Here $M$ is the diagonal mass matrix whose diagonal entries are the oscillator masses, ##V(q)## is the potential in which the oscillators move, and ##C(q)## is a positive definite damping matrix that determines the detailed friction behavior. it is easily seen that the total energy ##E=\frac12 M\dot q^2+ V(q)## is strictly decreasing as long as the velocity is nonzero, matching the experimental characteristics of friction. The case where ##C## is constant and ##V(q)_\frac12 q^TKq## with a constant stiffness matrix ##K## leads to the linear dynamical system ##M\ddot q+C\dot q +Kq=0##, which is the basis of most engineering calculations for friction in oscillating mechanical structures. More complex dissipative systems need a more complicated form of these equations, generalizing them to a Hamiltonian or Lagrangian framework.

In the quantum case, things are also a bit more complicated, though for different reasons. The correct high-level description of dissipative processes is given by a so-called Lindblad equation, a generalization of the quantum Liouville equation for conservative systems.
 
  • #5
I've done some elasticity calculations where we had to use a strain rate tensor to describe viscous friction in a deformable object. Simply adding a force proportional to velocity to the equations of motion of the volume elements of the object didn't work, as it would also damp the translational or rotational motion of the object as whole, too, even if it was supposed to be in vacuum.
 
  • #6
hilbert2 said:
I've done some elasticity calculations where we had to use a strain rate tensor to describe viscous friction in a deformable object. Simply adding a force proportional to velocity to the equations of motion of the volume elements of the object didn't work, as it would also damp the translational or rotational motion of the object as whole, too, even if it was supposed to be in vacuum.
Well, your article only describe ODEs, and I gave the most typical way to include dissipation, not the most general one. For PDEs such as in elasticity calculations, one needs schemes that also produce Navier-Stokes from Euler and can handle dissipative reaction-diffusion equations. None of these are of the form you described in your article. The book by Oettinger, Nonequilibrium Thermodynamics, might be a good starting point.
 
  • #7
If I make a velocity dependent potential energy ##V(x,\dot{x})=\frac{1}{2}kx^2 + \beta x \dot{x}## (implying that at any moment there's a linear potential that increases when going to the direction of motion), it seems that I get a velocity term in the equation of motion from the Lagrange equations. But then the generalized momentum ##p## seems to depend on ##x##, too, and the Hamiltonian function is not equal to the total energy, it seems to me. I admit I should have described this in more detail.
 
  • #8
hilbert2 said:
If I make a velocity dependent potential energy ##V(x,\dot{x})=\frac{1}{2}kx^2 + \beta x \dot{x}## (implying that at any moment there's a linear potential that increases when going to the direction of motion), it seems that I get a velocity term in the equation of motion from the Lagrange equations. But then the generalized momentum ##p## seems to depend on ##x##, too, and the Hamiltonian function is not equal to the total energy, it seems to me. I admit I should have described this in more detail.
From a variational principle for which the Hamiltonian is the energy, you never get dissipative equations. And if you change the Lagrangian such that it produces the right equations, and you don't introduce explicit time dependence (which would be unnatural in a system where there are no external forces) the Hamiltonian will typically be zeros. Thus the standard connection to the physics is lost.

The principles of Lagrange and Hamilton are taylored to the conservative case; this is why they are so prominent in the books. Dissipation is (in general, not in certain special cases) a much more complicated phenomenon. It is of thermodynamic origin (friction creates heat), and to get the correct equations one usually must go through a thermodynamic derivation from a microscopic description, or at least use equations of a form that are inspired from such a derivation.
 
  • Like
Likes Mentz114
  • #9
A. Neumaier said:
The principles of Lagrange and Hamilton are taylored to the conservative case; this is why they are so prominent in the books. Dissipation is (in general, not in certain special cases) a much more complicated phenomenon. It is of thermodynamic origin (friction creates heat), and to get the correct equations one usually must go through a thermodynamic derivation from a microscopic description, or at least use equations of a form that are inspired from such a derivation.

I'm familiar with articles like this, but haven't gone through the derivation myself. If there's too few oscillators in the heat bath, I guess there will also be Brownian motion-like fluctuation.
 
  • #10
Yes, this is the thermodynamic treatment There is fluctuation no matter how many bath oscillators you have. One couples a system to a heat bath (in the most typical case) and then eliminates the bath, gets complicated equations, makes a Markov approximation to eliminate the memory, is left with a stochastic differential equation, and if the noise is so small that it can be neglected one ends up with a dissipative deterministic equation. See, e.g., Chapters 15 and 16 of my online book.

For fluid flow and elasticity, there is no external heat bath; dissipation is energy lost into the high frequency modes, and one must start with a microscopic multiparticle system or quantum field theory.
 
  • #11
I tested the nonlinear Schrödinger equation that I mentioned in the insight as one way to produce effective damping in QM, and the correct way to form the damped oscillator TDSE seems to be

##i\hbar \frac{\partial \Psi}{\partial t} = -\frac{\hbar^2}{2m}\frac{\partial^2 \Psi}{\partial x^2} + \frac{1}{2}kx^2\Psi + \alpha Im\left(\int_{-\infty}^{x}\Psi (x',t)dx'\right)\Psi##,

where ##\alpha## is a positive constant. The ##Im## means imaginary part. This equation seems to conserve norm and if an initial state ##\Psi (x,0) = Ce^{-a(x-x_0 )^2}## is evolved with it, the time-dependent expectation value ##\left<x\right>## behaves like a damped classical oscillator.

I have a sample of a code (written in R) in my blog that calculates the evolution of a gaussian wavepacket in harmonic potential: https://physicscomputingblog.wordpress.com/2017/02/25/numerical-solution-of-pdes-part-4-schrodinger-equation/ .

A version of the code that adds the nonlinear damping looks like this:
Code:
library(graphics)                                    #load the graphics library needed for plotting

lx <- 6.0                                                        #length of the computational domain
lt <- 5.0                                        #length of the simulation time interval
nx <- 300                                      #number of discrete lattice points
nt <- 200                                       #number of timesteps
dx <- lx/nx                                    #length of one discrete lattice cell
dt <- lt/nt                                                    #length of timestep

V = c(1:nx)                                    #potential energies at discrete points

for(j in c(1:nx)) {
V[j] = as.complex(2*(j*dx-3)*(j*dx-3))                        #Harmonic oscillator potential with k=4
}

kappa1 = (1i)*dt/(2*dx*dx)                            #an element needed for the matrices                              
kappa2 <- c(1:nx)                                #another element

for(j in c(1:nx)) {
kappa2[j] <- as.complex(kappa1*2*dx*dx*V[j])
}

psi = as.complex(c(1:nx))                                                        #array for the wave function values

for(j in c(1:nx)) {
psi[j] = as.complex(exp(-(j*dx-2)*(j*dx-2)))                    #Gaussian initial wavefunction, displaced from equilibrium
}
  
xaxis <- c(1:nx)*dx                                #the x values corresponding to the discrete lattice points

for(m in c(1:nt)) {

A = matrix(nrow=nx,ncol=nx)                                #matrix for forward time evolution
B = matrix(nrow=nx,ncol=nx)                                #matrix for backward time evolution

for(j in c(1:nx)) {
kappa2[j] <- as.complex(kappa1*2*dx*dx*V[j])
for(k in c(1:j)) {
kappa2[j] = kappa2[j]+2*kappa1*2*dx*dx*dx*Im(psi[k])  # Add a nonlinear integral damping term to the potential
}
}

for(j in c(1:nx)) {
for(k in c(1:nx)) {
A[j,k]=0
B[j,k]=0
if(j==k) {
A[j,k] = 1 + 2*kappa1 + kappa2[j]                      
B[j,k] = 1 - 2*kappa1 - kappa2[j]
}
if((j==k+1) || (j==k-1)) {
A[j,k] = -kappa1                                  
B[j,k] = kappa1
}
}
}                                        #main time stepping loop

sol <- solve(A,B%*%psi)                         #solve the system of equations

for (l in c(1:nx)) {
psi[l] <- sol[l]
}

if(m %% 3 == 1) {                                                     #make plots of psi(x) on every third timestep
jpeg(file = paste("plot_",m,".jpg",sep=""))
plot(xaxis,abs(psi)^2, xlab="position (x)", ylab="Im(Psi)",ylim=c(-1.5,1.5))
title(paste("Abs(psi(x,t))^2 at t =",m*dt))
lines(xaxis,abs(psi)^2)
dev.off()
}
}

Just need to choose a short enough timestep and spatial step.
 
  • #12
hilbert2 said:
I tested the nonlinear Schrödinger equation that I mentioned in the insight as one way to produce effective damping in QM, and the correct way to form the damped oscillator TDSE seems to be
##i\hbar \frac{\partial \Psi}{\partial t} = -\frac{\hbar^2}{2m}\frac{\partial^2 \Psi}{\partial x^2} + \frac{1}{2}kx^2\Psi + \alpha Im\left(\int_{-\infty}^{x}\Psi (x',t)dx'\right)\Psi##,
But this is no longer a differential equation but an integro-differential equation with a memory term that depends on an infinitely long past. You cannot even solve it when you initially only have information at time $t=0$, as in most physical problems. So I don't understand how you can evolve your state at time $t=0$.

This might be ok for an academic computational project. But no engineer would be using such equations to study dissipative problems in fluid flow, say. Your approach is very ad hoc and specialized - it does not even cover such well-studied dissipative processes as the Navier-Stokes equations!
 
  • #13
The integration is only over ##x'##, and the variable ##t## in the integrand is set to the constant value of present time, so there is no memory. Therefore the solution can be approximately calculated by updating the potential ##V(x)## on each time step to correspond to the wavefunction ##\psi (x)## at that moment.

EDIT: Here's a video about fluid mechanics simulations where a nonlinear Schrödinger equation produces results equivalent to Navier-Stokes: .
 
  • #14
hilbert2 said:
the variable t in the integrand is set to the constant value of present time, so there is no memory.
Sorry for having misread your formula. Still, it is no longer a partial differential equation but a nonlocal dynamics that is very uncommon in practice.
 
  • #15
Energy dissipation in turbulent flow is certainly a problem of great interest to scientists, engineers, and mathematicians, but it is not the first path I would choose to study "damped motion." One natural analog of damped oscillator problems, which have generally been rendered artificially linear, that are studied in introductory mechanics would be to study problems that can be made artificially linear in fluid mechanics, whether compressible or not, or, say, magnetohydrodynamics, where Ohm's Law provides a dissipative knob that can be turned. Conflating the mechanics of linear viscous dissipation with the complicated mechanics of turbulent flow seems unhelpful as pedagogy.

Introduction of particle kinetics is another unnecessary distraction. Linear or nonlinear, deriving continuum equations by taking moments of the Boltzmann equation is tangential to the subject at hand. Standard approximations used by engineers, mathematicians, and even plasma physicists take care of the thermodynamics, which can also be made linear for studying, say, damped acoustic waves, so non-equilibrium thermodynamics would be interesting mostly if that and not "damped motion" were the subject at hand.

I'm not a solid mechanics wizard, so I don't know the details of how structural engineers do structural damping or how damping is modeled in seismic mechanics, but those also seem like better paths than wandering off into turbulent flow, nonequilibrium thermodynamics, particle kinetics, or other advanced topics.

I'm confident that there must be a need for studying "damped motion" or its analog in the context of quantum mechanics. I just don't know what it is.
 
  • Like
Likes Mentz114
  • #16
Robert M said:
I'm confident that there must be a need for studying "damped motion" or its analog in the context of quantum mechanics. I just don't know what it is.

The infrared and microwave spectral line broadening in solid and liquid phase samples can be modeled by using some kind of effective damping to represent the interaction of the rotating or vibrating molecules with neighboring atoms and molecules. This can be done by calculating the spectrum as a Fourier transform of an autocorrelation function ##\left<\psi(t_0 )|\psi (t)\right>##, as described here: http://home.uchicago.edu/~tokmakoff/TDQMS/Notes/6.1.-6.3_Absorption_Lineshape_2-25-09.pdf .
 
  • #17
The article has now been edited to clarify some things.
 
  • Like
Likes Greg Bernhardt
  • #18
hilbert2 said:
The infrared and microwave spectral line broadening in solid and liquid phase samples can be modeled by using some kind of effective damping to represent the interaction of the rotating or vibrating molecules with neighboring atoms and molecules. This can be done by calculating the spectrum as a Fourier transform of an autocorrelation function ##\left<\psi(t_0 )|\psi (t)\right>##, as described here: http://home.uchicago.edu/~tokmakoff/TDQMS/Notes/6.1.-6.3_Absorption_Lineshape_2-25-09.pdf .

Thanks. It occurred to me after I had posted my comment that there must be examples in solid state physics where it is generally thought to be appropriate to speak of damped motion, the first likely example of which that occurred to me was phonon damping. If I understand your offered article correctly, the "damping" is really relaxation from one state to another, with whatever energy is lost being carried off as radiation. I believe that's a fairly general characteristic of "damping" in quantum mechanics that distinguishes it from classical mechanics, where energy simply "disappears" (or is converted to heat).
 
  • #19
hilbert2 said:
The infrared and microwave spectral line broadening in solid and liquid phase samples can be modeled by using some kind of effective damping to represent the interaction of the rotating or vibrating molecules with neighboring atoms and molecules. This can be done by calculating the spectrum as a Fourier transform of an autocorrelation function ##\left<\psi(t_0 )|\psi (t)\right>##, as described here: http://home.uchicago.edu/~tokmakoff/TDQMS/Notes/6.1.-6.3_Absorption_Lineshape_2-25-09.pdf .

Line-broadening may, after all, be a better way to characterize the topic than anything to do with (say) opposing motion, as it generalizes naturally (as you have shown) to quantum mechanics. For many applications, the "Q" of resonant processes is more important for calculation than intuitive notions of frictional losses. A focus on line width also sidesteps a messy detour into thermodynamics and a necessarily complicated discussion of where energy is "lost" to.
 
  • #20
For holonomic systems, the inclusion of dissipation terms (and driving/exciting terms) in the equations of motion is not difficult if the system is formulated by means of Hamilton's Principle. This topic is developed at some length in Dynamics of Mechanical and Electromechanical Systems by Crandall, Karnopp, Kurtz, and Pridmore-Brown (McGraw-Hill, 1968). Provided the functional form of the frictional force can be expressed, it is rather automatic to include such terms in a virtual work term to be included in the variational indicator.
 
  • Like
Likes hilbert2
  • #21
I saw an announcement about a diploma thesis presentation (in the university I'm doing PhD studies in) today, which was on an interesting subject, namely active damping of oscillations (in contrast to passive friction forces). That seems to be based on monitoring a vibration and applying a force that's in anti-resonance with the motion. Here's a publication about a related thing: http://www.dspe.nl/files/MechatronicsPaper.pdf . This seems to be similar to my own idea of a damping term that is proportional to a momentum expectation value.
 
  • Like
Likes Robert M
  • #22
hilbert2 said:
I saw an announcement about a diploma thesis presentation (in the university I'm doing PhD studies in) today, which was on an interesting subject, namely active damping of oscillations (in contrast to passive friction forces).

Active dampers are used in skyscrapers to minimize wind-induced vibrations and seismic shaking. An active damper was part of solving the problem of windows popping out of their frames in the I M Pei-designed Hancock Tower in Boston.
 
  • Like
Likes Nidum

1. What is damped motion in classical and quantum mechanics?

Damped motion is a type of motion where the amplitude of a system decreases over time due to the presence of a dissipative force. In classical mechanics, this can be seen in systems such as a mass-spring-damper system. In quantum mechanics, it refers to the decay of a quantum state due to the system's interaction with its environment.

2. What is the difference between classical and quantum damped motion?

The main difference between classical and quantum damped motion lies in their underlying theories. Classical mechanics is based on Newton's laws of motion and describes the behavior of macroscopic objects, while quantum mechanics is based on the principles of quantum physics and describes the behavior of subatomic particles. In terms of damped motion, classical mechanics focuses on the decrease in amplitude of a system over time, while quantum mechanics focuses on the decay of a quantum state due to the system's interaction with its environment.

3. What are some examples of damped motion in classical and quantum systems?

In classical mechanics, examples of damped motion include a swinging pendulum, a car's suspension system, and a guitar string gradually losing its vibration. In quantum mechanics, examples include the decay of an excited atom emitting photons, the decay of a radioactive nucleus, and the decay of an unstable quantum state into a more stable state.

4. How does damped motion affect the behavior of a system?

Damped motion can significantly impact the behavior of a system. In classical mechanics, it can cause a system to eventually come to rest due to the dissipation of energy. In quantum mechanics, it can lead to the loss of coherence and information of a quantum state, which can have important implications for quantum computing and communication.

5. Can damped motion be reversed or controlled?

In classical mechanics, it is possible to reduce the effects of damped motion by minimizing the dissipation of energy through methods such as lubrication or reducing friction. In quantum mechanics, it is much more difficult to control damped motion as it is a fundamental property of quantum systems. However, researchers are exploring ways to reduce the effects of damped motion in quantum systems through quantum error correction techniques.

Similar threads

  • Quantum Physics
4
Replies
118
Views
12K
  • Quantum Physics
Replies
8
Views
2K
Replies
287
Views
24K
  • Quantum Physics
Replies
2
Views
2K
  • Quantum Physics
Replies
22
Views
6K
Replies
395
Views
20K
  • Quantum Physics
Replies
15
Views
2K
Replies
9
Views
2K
Replies
1
Views
1K
Replies
82
Views
8K
Back
Top