Linear response function

A linear response function describes the input-output relationship of a signal transducer, such as a radio turning electromagnetic waves into music or a neuron turning synaptic input into a response. Because of its many applications in information theory, physics and engineering there exist alternative names for specific linear response functions such as susceptibility, impulse response or impedance; see also transfer function. The concept of a Green's function or fundamental solution of an ordinary differential equation is closely related.

Mathematical definition

Denote the input of a system by h ( t ) {\displaystyle h(t)} (e.g. a force), and the response of the system by x ( t ) {\displaystyle x(t)} (e.g. a position). Generally, the value of x ( t ) {\displaystyle x(t)} will depend not only on the present value of h ( t ) {\displaystyle h(t)} , but also on past values. Approximately x ( t ) {\displaystyle x(t)} is a weighted sum of the previous values of h ( t ) {\displaystyle h(t')} , with the weights given by the linear response function χ ( t t ) {\displaystyle \chi (t-t')} :

x ( t ) = t d t χ ( t t ) h ( t ) + . {\displaystyle x(t)=\int _{-\infty }^{t}dt'\,\chi (t-t')h(t')+\cdots \,.}

The explicit term on the right-hand side is the leading order term of a Volterra expansion for the full nonlinear response. If the system in question is highly non-linear, higher order terms in the expansion, denoted by the dots, become important and the signal transducer cannot adequately be described just by its linear response function.

The complex-valued Fourier transform χ ~ ( ω ) {\displaystyle {\tilde {\chi }}(\omega )} of the linear response function is very useful as it describes the output of the system if the input is a sine wave h ( t ) = h 0 sin ( ω t ) {\displaystyle h(t)=h_{0}\sin(\omega t)} with frequency ω {\displaystyle \omega } . The output reads

x ( t ) = | χ ~ ( ω ) | h 0 sin ( ω t + arg χ ~ ( ω ) ) , {\displaystyle x(t)=\left|{\tilde {\chi }}(\omega )\right|h_{0}\sin(\omega t+\arg {\tilde {\chi }}(\omega ))\,,}

with amplitude gain | χ ~ ( ω ) | {\displaystyle |{\tilde {\chi }}(\omega )|} and phase shift arg χ ~ ( ω ) {\displaystyle \arg {\tilde {\chi }}(\omega )} .

Example

Consider a damped harmonic oscillator with input given by an external driving force h ( t ) {\displaystyle h(t)} ,

x ¨ ( t ) + γ x ˙ ( t ) + ω 0 2 x ( t ) = h ( t ) . {\displaystyle {\ddot {x}}(t)+\gamma {\dot {x}}(t)+\omega _{0}^{2}x(t)=h(t).}

The complex-valued Fourier transform of the linear response function is given by

χ ~ ( ω ) = x ~ ( ω ) h ~ ( ω ) = 1 ω 0 2 ω 2 + i γ ω . {\displaystyle {\tilde {\chi }}(\omega )={\frac {{\tilde {x}}(\omega )}{{\tilde {h}}(\omega )}}={\frac {1}{\omega _{0}^{2}-\omega ^{2}+i\gamma \omega }}.}

The amplitude gain is given by the magnitude of the complex number χ ~ ( ω ) , {\displaystyle {\tilde {\chi }}(\omega ),} and the phase shift by the arctan of the imaginary part of the function divided by the real one.

From this representation, we see that for small γ {\displaystyle \gamma } the Fourier transform χ ~ ( ω ) {\displaystyle {\tilde {\chi }}(\omega )} of the linear response function yields a pronounced maximum ("Resonance") at the frequency ω ω 0 {\displaystyle \omega \approx \omega _{0}} . The linear response function for a harmonic oscillator is mathematically identical to that of an RLC circuit. The width of the maximum, Δ ω , {\displaystyle \Delta \omega ,} typically is much smaller than ω 0 , {\displaystyle \omega _{0},} so that the Quality factor Q := ω 0 / Δ ω {\displaystyle Q:=\omega _{0}/\Delta \omega } can be extremely large.

Kubo formula

The exposition of linear response theory, in the context of quantum statistics, can be found in a paper by Ryogo Kubo.[1] This defines particularly the Kubo formula, which considers the general case that the "force" h(t) is a perturbation of the basic operator of the system, the Hamiltonian, H ^ 0 H ^ 0 h ( t ) B ^ ( t ) {\displaystyle {\hat {H}}_{0}\to {\hat {H}}_{0}-h(t'){\hat {B}}(t')} where B ^ {\displaystyle {\hat {B}}} corresponds to a measurable quantity as input, while the output x(t) is the perturbation of the thermal expectation of another measurable quantity A ^ ( t ) {\displaystyle {\hat {A}}(t)} . The Kubo formula then defines the quantum-statistical calculation of the susceptibility χ ( t t ) {\displaystyle \chi (t-t')} by a general formula involving only the mentioned operators.

As a consequence of the principle of causality the complex-valued function χ ~ ( ω ) {\displaystyle {\tilde {\chi }}(\omega )} has poles only in the lower half-plane. This leads to the Kramers–Kronig relations, which relates the real and the imaginary parts of χ ~ ( ω ) {\displaystyle {\tilde {\chi }}(\omega )} by integration. The simplest example is once more the damped harmonic oscillator.[2]

See also

References

  1. ^ Kubo, R., Statistical Mechanical Theory of Irreversible Processes I, Journal of the Physical Society of Japan, vol. 12, pp. 570–586 (1957).
  2. ^ De Clozeaux,Linear Response Theory, in: E. Antončik et al., Theory of condensed matter, IAEA Vienna, 1968

External links

  • Linear Response Functions in Eva Pavarini, Erik Koch, Dieter Vollhardt, and Alexander Lichtenstein (eds.): DMFT at 25: Infinite Dimensions, Verlag des Forschungszentrum Jülich, 2014 ISBN 978-3-89336-953-9