Jump to content

User:Emitabsorb/sandbox

From Wikipedia, the free encyclopedia

Derivation[edit]

Quantum version[edit]

The fluctuation-dissipation theorem relates the correlation function of the observable of interest (a measure of fluctuation) to the imaginary part of the response function (a measure of dissipation), in the frequency domain. A link between these quantities can be found through the so-called Kubo formula [1]

which follows, under the assumptions of the linear response theory, from the time evolution of the ensemble average of the observable in the presence of a perturbing source. The Kubo formula allows us to write the imaginary part of the response function as

In the canonical ensemble, the second term can be re-expressed as

where in the second equality we re-positioned using the cylic property of trace (in this step we have also assumed that the operator is bosonic, i.e. does not introduce a sign change under permutation). Next, in the third equality, we inserted next to the trace and interpreted as a time evolution operator with imaginary time interval . We can then Fourier transform the imaginary part of the response function above to arrive at the quantum fluctuation-dissipation relation [2]

where is the Fourier transform of and is the Bose-Einstein distribution function. The "" term can be thought of as due to quantum fluctuations. At high enough temperatures, , i.e. the quantum contribution is negligible, and we recover the classical version.

Examples in detail[edit]

The fluctuation–dissipation theorem is a general result of statistical thermodynamics that quantifies the relation between the fluctuations in a system that obeys detailed balance and the response of the system to applied perturbations.

Brownian motion[edit]

For example, Albert Einstein noted in his 1905 paper on Brownian motion that the same random forces that cause the erratic motion of a particle in Brownian motion would also cause drag if the particle were pulled through the fluid. In other words, the fluctuation of the particle at rest has the same origin as the dissipative frictional force one must do work against, if one tries to perturb the system in a particular direction. From this observation Einstein was able to use statistical mechanics to derive the Einstein–Smoluchowski relation [3]

which connects the diffusion constant D and the particle mobility μ, the ratio of the particle's terminal drift velocity to an applied force. kB is the Boltzmann constant, and T is the absolute temperature.


Mathematically, the motion of a particle of mass undergoing a classical Brownian motion in one dimension is governed by the Langevin equation [3]

where is the velocity of the particle, is a frictional force and is a random force with the property , both of these forces are due to the particle's interaction with a thermal bath. In this example, the observable of interest is the velocity of the particle and to demonstrate the fluctuation-dissipation theorem, we determine both sides of the relation separately. The response function can be found immediately by Fourier transforming the Langevin equation

from which we can extract its imaginary part

This will be the dissipation side of the fluctuation-dissipation relation. Multiplying the Langevin equation by and taking the ensemble average, we find

The last term drops because the random force is independent of . Solving the remaining equation and using time-translational invariance to shift , we obtain

Fourier transforming this correlation function and applying the equipartition theorem for a single particle in one dimension

we obtain the fluctuation side of the fluctuation-dissipation relation

demonstrating the validity of the fluctuation-dissipation theorem.

Thermal noise in a resistor[edit]

In 1928, John B. Johnson discovered and Harry Nyquist explained Johnson–Nyquist noise. With no applied current, the mean-square voltage depends on the resistance , , and the bandwidth over which the voltage is measured [3]:

A simple circuit for illustrating Johnson-Nyquist thermal noise in a resistor.

This observation can be understood through the lens of the fluctuation-dissipation theorem. Take, for example, a simple circuit consisting of a resistor with a resistance and a capacitor with a small capacitance . Kirchhoff's law yields

and so the response function for this circuit is

In the low-frequency limit , its imaginary part is simply

which then can be linked to the auto-correlation function of the voltage via the fluctuation-dissipation theorem

The Johnson-Nyquist voltage noise was observed within a small frequency bandwidth centered around . Hence

Violations in glassy systems[edit]

While the fluctuation–dissipation theorem provides a general relation between the response of systems obeying detailed balance, when detailed balance is violated comparison of fluctuations to dissipation is more complex. Below the so called glass temperature , glassy systems are not equilibrated, and slowly approach their equilibrium state. This slow approach to equilibrium is synonymous with the violation of detailed balance. Thus these systems require large time-scales to be studied while they slowly move toward equilibrium.


File:Violation of FDT1.png
Violation of the fluctuation-dissipation theorem (FDT) in the Edwards-Anderson system with temperature , magnetic susceptibility , and spin-temporal correlation function . Figure replotted from [4].

To study the violation of the fluctuation-dissipation relation in glassy systems, particularly spin glasses, Ref. [4] performed numerical simulations of macroscopic systems (i.e. large compared to their correlation lengths) described by the three-dimensional Edwards-Anderson model using supercomputers. In their simulations, the system is initially prepared at a high temperature, rapidly cooled to a temperature below the glass temperature , and left to equilibrate for a very long time under a magnetic field . Then, at a later time , two dynamical observables are probed, namely the response function

and the spin-temporal correlation function

where is the spin living on the node of the cubic lattice of volume , and is the magnetization density. The fluctuation-dissipation relation in this system can be written in terms of these observables as

Their results confirm the expectation that as the system is left to equilibrate for longer times, the fluctuation-dissipation relation is closer to be satisfied.


In the mid-1990s, in the study of dynamics of spin glass models, a generalization of the fluctuation–dissipation theorem was discovered [5] that holds for asymptotic non-stationary states, where the temperature appearing in the equilibrium relation is substituted by an effective temperature with a non-trivial dependence on the time scales. This relation is proposed to hold in glassy systems beyond the models for which it was initially found.

Applications[edit]

Early Universe Cosmology[edit]

In the paradigm of the widely successful Big Bang theory, the Universe started out at an extremely high temperature and gradually cooled down as it expanded. In the process, the temperature of the Universe occasionally went pass various important energy scales such as the Quantum Chromodynamics scale, triggering various phase transitions driving (part of) the Universe out of thermal equilibrium. The evolution of the Universe as it relaxed back to thermal equilibrium following these cosmological phase transitions is governed by some dissipation which inevitably comes with the accompanying fluctuations [6][7].

  1. ^ "The fluctuation-dissipation theorem". doi:10.1088/0034-4885/29/1/306. {{cite journal}}: Cite journal requires |journal= (help)
  2. ^ "Fundamental aspects of quantum Brownian motion". doi:10.1063/1.1853631. {{cite journal}}: Cite journal requires |journal= (help)
  3. ^ a b c Blundell, Stephen J.; Blundell, Katherine M. (2009). Concepts in thermal physics. OUP Oxford.
  4. ^ a b "A statics-dynamics equivalence through the fluctuation–dissipation ratio provides a window into the spin-glass phase from nonequilibrium measurements". doi:10.1073/pnas.1621242114. {{cite journal}}: Cite journal requires |journal= (help)
  5. ^ "Analytical solution of the off-equilibrium dynamics of a long-range spin-glass model". doi:10.1103/PhysRevLett.71.173. {{cite journal}}: Cite journal requires |journal= (help)
  6. ^ "One-loop fluctuation-dissipation formula for bubble-wall velocity". doi:10.1103/PhysRevD.48.1539. {{cite journal}}: Cite journal requires |journal= (help)
  7. ^ "Fluctuation-dissipation dynamics of cosmological scalar fields". doi:10.1103/PhysRevD.91.083540. {{cite journal}}: Cite journal requires |journal= (help)