Linear Response in Classical Physics

Size: px
Start display at page:

Download "Linear Response in Classical Physics"

Transcription

1 Linear Response in Classical Physics Wouter G. Ellenbroek Technische Universiteit Eindhoven Notes (section ) by Fred MacKintosh (Vrije Universiteit), used previously in the 11 DRSTP course 1 Introduction The theory of linear response deals with the question how the measurable properties of a physical system change when it is slightly taken away from its equilibrium state. Traditionally, the term refers to linear response of thermal systems to slight deviations from thermodynamic equilibrium. In the first part of this course, we will deal with the question how this response is related to the equilibrium thermal fluctuations in the system. The notes for this first part have been provided by Fred MacKintosh, who taught a similar course in a DRSTP school a few years ago. In the second part, we will consider an athermal system that is in mechanical equilibrium, but not in thermodynamic equilibrium, and study its (linear) response to mechanical perturbations. The course will be followed up with a course on quantum linear response theory, taught by Rembert Duine. Fluctuations and Linear Response.1 Brownian Motion: Random Walks in 1d We begin with arguably the simplest problem illustrating the effect of fluctuations, e.g., in a fluid: Brownian motion. This term derives from the botanist Robert Brown who observed random motions of pollen grains in a fluid in 188. This kind of motion is general to sufficiently small particles suspended in a fluid. By observing different kinds of particles, he concluded that the motion was not dependent on the type of particles. We now know that this motion is due to the ever-present thermal fluctuations in the fluid itself. In a series of papers beginning in 195, Einstein gave a theoretical basis for Brownian motion. About the same time, Smoluchowski derived the same results in a somewhat different way. Here, we derive the properties of what is usually referred to as the random walk problem for a single particle. Despite the seeming simplicity, this constitutes a rich example that will help to illustrate many of the principles of fluctu- 1

2 ation about equilibrium generally. Examples of such random walks include the Brownian motion of a particle in a fluid, as well as the random conformations of long polymer molecules under some conditions. In the first of these examples, the essential physical aspects of this problem are: the existence of randomly fluctuating forces pushing the particle in random directions, the viscous drag of the particle in the fluid, and the inertia of the particle. As we shall see later, the the drag of the particle (or equivalently, its mobility when subject to an external force) is fundamentally related to the strength of the random forces that it experiences in the fluid. We ll see later a statement of this deep connection between fluctuations and dissipation in the form of the so-called Fluctuation- Dissipation Theorem, and it is largely with this in mind that we shall examine the basic problem of Brownian motion in such great detail. We first look at Brownian motion in just one spatial dimension. Consider a particle that can move just left (L) and right (R) along a line by steps of equal length l. Steps to the left and right occur randomly, with equal probability p = 1/. Because the probability to jump to the left and to the right are equal, the average displacement of a particle during the i-th step is zero: x i = 1 (+l) + 1 ( l) =. (1) However, the mean squared displacement is non zero: x i = 1 (+l) + 1 ( l) = l. () Next consider average displacement after N steps: X = N x i =. (3) i=1 The mean squared displacement after N steps is: X = N i=1 j=1 N x i x j. (4) The jump directions on different steps are not correlated, therefore x i x j = for i j. (5) For i = j, we have x i x j = l. If we insert this in the expression for X we get N N X = l δ ij. (6) where δ ij is a Kronecker delta: i=1 j=1 δ ij 1 for i = j = for i j.

3 Hence: X = Nl. (7) If the time interval between successive jumps of the particle is δt, then the number of jumps in a time interval t is equal to t/δt and X (t) = l t/δt. (8) In other words, the mean squared displacement increases linearly with time. We should compare this with the ballistic motion of a particle with a velocity v, in which case X(t) = vt, and hence X (t) = v t. For the simple case of one-dimensional Brownian motion we can compute not just the mean-squared displacement of a particle, but the complete probability distribution to find a particle at a distance x from its original position, after N steps. We denote by N L and N R the number of steps that occur to the left and right. The total number of steps is then N = N L + N R, while the net length of the walk (to the right) is x = l (N R N L ). The number of such walks total steps N, of which N L steps are to the left and N R to the right is Ω (x, N) = N! N L!N R!. (9) Then, the probability of a such a walk of net length x is P (x, N) = N! N L!N R! ( ) N 1 = ( N x/l N! )! ( N+x/l By using Stirling s approximation n! = πnn n e n, we find that )! ( ) N 1. (1) P (x, N) N N e N R+N L πn = ( ) [N+x/l]/ ( ) [N x/l]/ N e N N+x/l N x/l πnr πnl πn = (1 + x/nl) [ N+x/l ] (1 x/nl) [ N x/l ] π(n + x/l) π(n x/l) = ( x/l (1 (x/nl) ) N/ (1+x/Nl) (1 x/nl)) Nπ(1 x /(Nl) ) Now we make use of the fact that and that ln (1 (x/nl) ) N/ exp( x /(Nl )) ( ) x/l (1 + x/nl) (x/l) (x/(nl) + x/(nl)) = x /(Nl ) (1 x/nl) 3

4 and hence We can then write ( ) x/l (1 + x/nl) = exp(x /(Nl )). (1 x/nl) P (x, N) 1 = πn exp( x /(Nl) ) exp(x /(Nl )) 1 x /(Nl) πn exp( x /(Nl )) (11) Of course, it comes as no surprise that this probability distribution is symmetric and that the most likely walk returns to the origin (x = ). Thus, the average x =. The mean-square displacement for such random walks, however, is nonzero. This is most easily calculated by treating P (x, N) above as a continuous distribution in the variable x. In fact, for small enough x/l, the distribution above varies slowly over the discrete values of x that are possible for a given large value of N. Thus, x x e x /(Nl ) dx = e x /(Nl ) dx = Nl. (1) Here, we have assumed that the particle moves with independent random steps of length l. If we assume further that these steps occur at equally-spaced time intervals δt, then t = Nδt and x = tl /δt. (13) This behavior x t or x t is characteristic of diffusion. Before leaving this simple one-dimensional random walk problem, we can make another observation. This problem is instructive for other physical situations, such as the conformations that a long polymer chain can make. The resulting conformations are referred to as Gaussian. We ignore the fact that the polymer chains cannot overlap in space. (Although an unphysical microscopic assumption, there are real situations where polymer conformations are nevertheless Gaussian.) Then, Ω(x, N) above represents the number of such conformations, subject to a given end-to-end separation x. Thus, by the fundamental assumption of statistical mechanics (that microscopic states of the same energy are equally likely), the probability of having a given value of x is proportional to Ω(x, N), the number of such conformations or states of the system. Since, for this simple problem, there are N states in all (N independent steps, with choices at each), the normalized probability is just P (x, N) above.. Diffusion One can also study Brownian motion from the point of view of diffusion. Here, we denote the density of (identical) particles suspended in the fluid by a concentration variable n( r, t), which is a function of both position r and time t. 4

5 We also denote by j( r, t) = n( r, t) v( r, t) the current density of particles at r and time t, where v is the velocity. In thermal equilibrium, and in the absence of other forces, the average density n( r, t) will be constant. Furthermore, by the Brownian motion of particles in the fluid, an initial inhomogeneity in the density will relax over time toward this average, uniform density. Thus, spatial variations (gradients) in n result in net particle motion, i.e., a particle current. For small variations in n, we expect that the corresponding current j is small. Thus, it is natural to assume a linear relationship: j = D n, (14) which is known as Fick s law. As we shall see, the minus sign agrees with our physical intuition that Brownian motion will lead over time to a uniform particle density. A positive sign here would lead to unphysical results. Thus, we see immediately that Brownian motion or diffusion is essentially dissipative and irreversible. The constant D here is known as the diffusion constant, which has units of square length per time. There is one more physical aspect to Brownian motion that we have neglected in this treatment so far, namely the fact that particles are conserved. If no particles are created or destroyed while they fluctuate around in the fluid, then the particle current and density are related by conservation n( r, t) t + j( r, t) =, which is simply a mathematical expression of the fact that a net particle current into (out of) any volume results in an increase (decrease) in the number of particles in that region. This, together with Fick s law leads to the diffusion equation, n( r, t) = D n( r, t). t A particular solution to this equation is n( r, t) = N e r 3/ (4πDt) /(4Dt). (15) This is, in fact, the solution 1 to the initial condition that N particles are placed at the origin at time t =, i.e., n( r, ) = N δ( r). This result is for a number N of particles in the fluid. Provided that these do not interact with each other, which is valid if the density is not too high, then we could have equivalently described things in terms of the probability density P ( r, t) for a single particle, which is given by Eq. (15) with N = 1. We then find that r(t) = ; r (t) = r P ( r, t)d 3 r = 6Dt t, 1 See Appendix A. 5

6 which is in agreement with our earlier result for the one-dimensional random walk: x(t) = ; x (t) = tl /δt = Dt t, where D = l /(δt). The general result depends on the number of dimensions d: r (t) = ddt. Even without solving the diffusion equation explicitly, we can relate x (t) to D. We do this as follows. We start with the one-dimensional diffusion equation for the probability density P (x, t): P (x, t) t = D P (x, t) x. We now multiply both sides of this equation by x and integrate over x: P (x, t) dx x = t We can write the left hand side of this equation as P (x, t) dx x = t t dx x D P (x, t) x. dx x P (x, t) = x (t) t The right-hand side can be computed by partial integration (making use of the fact that P (x, t) and its derivative vanish at x = ± : dx x D P (x, t) x P (x, t) = dx xd = +D dx P (x, t). x But as P (x, t) is normalized to one, the result is simply D and therefore: This we can integrate to yield x (t) t = D (16) x (t) = Dt where we have used the fact that at t =, the probability distribution is a δ-function. We can write similar expressions for y (t) and z (t). Adding these, we obtain r (t) = x (t) + y (t) + z (t) = 6Dt. In reality, for the example of a particle suspended in a fluid, we expect that the kicks that the particle experiences in the fluid are random. Even so, since the particle has a mass, there will be some effect of its inertia. So, we expect that the motion is not truly random, but will be correlated for short times. For such short times, the particle will move ballistically, and we expect that the displacement grows linearly with time. Thus, we expect that x t, rather than x t 6

7 for short enough times. Actually, this must be true. We can already see that there is a problem with the Eq. (13), because it would imply that x t, which leads to an unphysical divergence of the velocity at short times. Thus, only for times long compared with some microscopic correlation time (which we can be more quantitative about below) will diffusion, characterized by Eq. (13), be valid..3 Velocity correlation function x(t) is the distance that a particle has traveled in the x-direction in a time interval t. If we denote the instantaneous velocity in the x-direction by v x (t), then we can write x(t) = dt v x (t ). We can then write the mean-squared displacement as x (t) = dt dt v x (t )v x (t ). The quantity v x (t )v x (t ) is an example of a time correlation function - in this case, the velocity auto-correlation function (VACF). This quantity measures the degree of correlation between the velocity of a particle at two different times t and t. When t = t, the value of the of the velocity correlation function is simply v. At sufficiently long times, when the particle has lost all memory of its original velocity, v x (t )v x (t ) = v x (t ) v x (t ) = (because the average velocity of a particle in equilibrium is zero). Time correlation functions, such as the VACF are properties of the system in equilibrium. Their value can therefore never depend on the times t and t individually, but only on the difference. This is so because the properties of a system in equilibrium are time invariant, i.e.: If we choose τ = t, we get v x (t )v x (t ) = v x (t + τ)v x (t + τ) ( τ) v x (t )v x (t ) = v x (t t )v x (), which shows that the VACF depends only on the difference t t. Another important property of (classical) time correlation functions is that we can permute v x (t ) and v x (t ): v x (t )v x (t ) = v x (t )v x (t ). Because of this symmetry, we can write dt dt v x (t )v x (t ) = dt dt v x (t )v x (t ). 7

8 We make use of this result in evaluating x (t) t = t = = dt dt v x (t)v x (t ) dτ v x (τ)v x (), dt v x (t )v x (t ) where we have defined τ t t. If we insert the above result in Eqn. 16, we obtain D = dτ v x (τ)v x (). (17) This equation is only valid for times much longer than the decay time of correlations in the velocity, and hence we may replace the upper limit in the integral by +. The expression that results is D =.4 Langevin Theory dτ v x (τ)v x (). (18) In order to see at a deeper level the connection between the random fluctuations in the fluid and the dissipative nature of the resulting particle motion, we turn now to an approach suggested by Langevin. This will also solve a problem noted above, namely that there must be a failure of our prior analysis of Brownian motion at short times. Our approach here will be to more directly account for the fluctuating forces acting on the particles by the surrounding fluid. We shall see how thermal equilibrium (e.g., the expected uniform spatial distribution of particles) can be established by these fluctuations. We begin with the simple equation of motion satisfied by the particle of mass m m d v dt = F(t), where F is the force acting on the particle. We might be tempted at this point to assume that this force is random, and independent of the particle motion. After all, these forces are due to the fluctuating solvent, right? In fact, there are two physically distinct parts to this force: (i) the viscous drag on the particle, which depends on v and (ii) the randomly fluctuating forces ζ(t). We assume here that the first of these can be described by the Stokes formula for the drag on a spherical particle in a viscous fluid: 6πηa v, where η is the fluid viscosity and a is the radius of the sphere. This is valid for small particles and small velocities. It is important to note that this force 8

9 is random and has zero mean, only to the extent that the particle velocity is random with zero mean. Thus, we can say that, in the absence of external forces, it has a zero time average. It is important also to distinguish another kind of average that we have been a bit sloppy about so far: the ensemble average. This represents an average over a large number of similarly prepared systems (an ensemble). We can consider, for instance, the motion of a particle with some initial condition, say velocity v() at time t =. We expect that this initial velocity will change over time, and that it will become uncorrelated with the velocity at much later times t. Let τ denote this correlation time, after which memory of the initial conditions is lost. As a specific example, if this initial velocity were large compared with the expected thermal velocities, we would expect that it would decay toward zero for times large compared with τ. But, the precise way in which it will do this depends on the details of the fluctuating force ζ due to the fluid. We are not interested in the exact trajectory that an individual particle makes due to each and every collision with individual fluid molecules. Rather, we want to know what to expect for the way the particle velocity decays in a typical sample of a certain kind of solvent (e.g., with known viscosity η) at a certain temperature. Thus, we want to know how the particle velocity behaves on the average, i.e., the ensemble average. However, the ensemble average of the random force is zero ζ(t) =, while the ensemble average of the particle velocity is not zero, at least for times smaller than τ. Mathematically, we could say that the ensemble corresponds to a large set of possible functions ζ(t) consistent with a given fluid at a given temperature. The properties of these functions are yet to be determined, but that the average ζ(t) over this set is zero for all times t is clear by construction. At any time t, ensemble average quantities can be obtained by averaging over the possible functions ζ(t). We rewrite the equation of motion as follows m d v(t) = 6πηa v(t) + ζ(t). dt (19) If we take the average of this over the ensemble, we obtain which implies that m d v(t) dt = 6πηa v(t), v(t) = v()e t/τ, where we can now identify τ = m/(6πηa) as the relaxation time for effects of the initial condition to die out. Here again the essentially dissipative nature of this result is evident by the fact that this result cannot be reversed in time, although the precise connection of this irreversibility to the fluctuating forces ζ is still unclear. This is what we want to establish next. We can take the scalar product of (19) with the instantaneous position r(t). In so doing, we note that r d v dt = d dt ( r v) v. The result is m d dt r v = 6πηa r v + m v. () 9

10 We have used the fact that r ζ =. This last point is not as obvious as it may appear at first sight. Although r ζ =, it can be shown that v ζ. In the equation above, if we further assume that the Brownian particle has attained thermal equilibrium, then we can replace v by the equipartition value 3kT/m. (The kinetic energy mv / of the particle involves three independent degrees of freedom, in which the energy is quadratic. Classical Statistical Mechanics then predicts that this energy is 3 kt on average.) The resulting simple differential equation can be solved: r v = Ce t/τ + kt πηa. But, the constant C can be determined by the initial condition, which we take to be that r = at t = : r v = kt ( ) 1 e t/τ. πηa We can integrate this one more time, to find that r = kt [ ( )] t τ 1 e t/τ, πηa d since r v = 1 dt r. At long times, for which t τ, we find again that r = 6Dt. Furthermore, we can now identify the diffusion coefficient with the dissipative mechanism of the fluid viscosity: D = kt 6πηa. (1) We also find, as expected, that for short times, r t, thus resolving the problem with our earlier analysis for short times. More generally, there is a fundamental relationship between the particle diffusion constant D and the mobility µ or response of the particle to external forces. If we apply an external force f to the particle, say due to gravity, then the resulting drift velocity is v = µ f. This defines what we mean by the mobility. For our Stokes sphere in the viscous fluid, the full equation of motion becomes m d v(t) dt Thus, the mobility in this case is µ = 1 6πηa, and = 6πηa v(t) + f + ζ(t). () D = µkt. (3) This is known as the Einstein relation, and is more general than this particular problem of a sphere in a viscous fluid. It is also an example of the fundamental 1

11 relationship that exists between a dissipative response and fluctuations. In this case, the response is that of a particle to an external field. This response is characterized by the mobility µ. The fluctuating quantity here is the particle position r(t). There is another way to derive the Einstein relation that makes the relation between the mobility µ and the diffusion constant D more obvious. Consider solution in a closed volume V. If there is a concentration gradient of the dissolved species, this will result in a diffusion current j. The diffusion is equal to the number density ρ of the dissolved species, multiplied by the average velocity of these particles. j = ρ v. Now suppose that the dissolve particles are subject to an external potential U(x). If this potential is not constant, there will be a net force acting on the particles: f x = U(x), x where we have assumed that the potential ia a function of x only. The average velocity of a particle due to this force is v x = µf x = µ U(x) x As the particles move under the influence of this force, the density will become inhomogeneous. But once the density is not constant, there will also be a diffusion current. When the system reaches equilibrium, the diffusion current is exactly balanced by the current due to the external force, i.e.:. = ρ v x = ρ(x)µ U x D ρ(x) x. (4) But we also know that, in equilibrium, the probability to find a particle at a position x must be proportional to the Boltzmann factor exp( U(x)/kT ): ρ(x) = ρ exp( U(x)/kT ). If we insert this expression in Eqn. 4, we get This equation is satisfied for all x if = ρ v x = ρ(x)µ U x + D kt ρ(x) U(x). (5) x D = µkt, that is, the Einstein relation. We can actually establish a more detailed connection between dissipation (specifically, the drag coefficient α = 6πηa here) and the fluctuating force ζ(t). To do this, we rewrite the equation of motion, dividing out by the particle mass d dt v(t) = 1 τ v(t) + 1 m ζ(t). (6) 11

12 This, we rearrange slightly and multiply by e t/τ : e t/τ d dt v(t) + 1 τ et/τ v(t) = d ( ) e t/τ v(t) = e t/τ 1 ζ(t). dt m (7) We now change t to t and integrate from to t, with the result that v(t) = v()e t/τ + e t/τ 1 m e t /τ ζ(t )dt. (8) Again, we see that if we take the ensemble average of this equation, we obtain v(t) = v()e t/τ, as we found before. This is because ζ(t ) = for any t. If we square both sides of (8) and then take the ensemble average, however, we find that v = v() e t/τ + e t/τ 1 m e (t +t )/τ ζ(t ) ζ(t ) dt dt. (9) Again, a term had been dropped because ζ =. The correlation function K(t, t ) = ζ(t) ζ(t ), however, is non-zero in general. This is where the ugly details of the fluctuating molecules in the fluid begin to appear directly in our analysis. We finally have a measurable quantity that would appear to depend directly on the random forces ζ. These appear through the correlation function K(t, t ) = ζ(t) ζ(t ), which has several general properties that we can identify. First of all, so long as we are looking at an equilibrium system that is not evolving in any macroscopic sense in time, then this function can only depend on the time interval s = t t : K(t, t ) = K(s). Furthermore, provided that we confine our attention to classical variables (as opposed to possibly non-commuting quantum mechanical operators), we can reverse the order of ζ(t) and ζ(t ), which tells us that K(s) is symmetric, i.e., that it depends only on s. On simple physical grounds, we also expect that K(s) decreases as the time interval s increases, since we expect the force ζ(t) to lose memory of its earlier values. More precisely, we expect that ζ(t) and ζ(t ) become decorrelated as t t grows, i.e., that K(s) = ζ(s) ζ() ζ(s) ζ() = as s grows. Furthermore, we can see that K(s) attains its maximum value at s =. In fact, since [ ζ(s) ± ζ() ] = ζ (s) + ζ () ± ζ(s) ζ() = (K() ± K(s)), the function K(s) is bounded by the limits K() and K(). In particular, K(s) K(). This all leads us to the inevitable conclusion that K(s) = ζ(s) ζ() is a function peaked at s =, from which it decays rapidly toward zero in a short 1

13 time τ, which is determined by the fast dynamics of the individual molecules of the fluid surrounding the Brownian particle. We expect a very short time scale for this, of order the time between collisions with the individual molecules. More generally, it should be the time scale that characterizes the rate with which the degrees of freedom responsible for the random forces ζ relax to internal equilibrium after a displacement of the particle. In fact, for all practical purposes, this correlation function decays so quickly away from s = that we may take it to be proportional to the delta function, that is that it can be taken to be non-zero only for s =, but with finite integral over s: C = K(s)ds. Returning to the problem that led us to consider this function K(s) in the first place, namely the evaluation of the mean-square velocity, we note that the double integral in (9) is appreciably non-zero only for t = t. Thus, v = v() e t/τ + e t/τ C m e t /τ dt = v() e t/τ + e t/τ C m τ ( e t/τ 1 where we have made the approximation suggested above: K(s) = Cδ(s). ), (3) In particular, however, Eq. 3 is true for times t large compared with τ, at which point all memory of the initial conditions (e.g., v()) should have been forgotten. Then, the ensemble average v = 3kT/m, consistent with the Equipartition Theorem. The only way this can be true is if Cτ m = 3kT, i.e., that the drag coefficient α is fundamentally related to the random forces ζ: α = 1 6kT ζ(s) ζ() ds. (31) The above result can be obtained in a few lines, using the techniques described in Appendix B (see, in particular, section B.1). The relation between the drag coefficient and the random forces acting on a particle is special case of what we will later learn generally as the Fluctuation-Dissipation theorem. For now, we note that the drag that a particle experiences in the fluid is due, at a microscopic level, to the randomly fluctuating forces that the particle feels at finite temperature. The random forces are at the origin of macroscopic dissipation. We began this analysis of the motion of a particle in a dissipative medium by considering the equation of motion m d v(t) dt = α v(t) + f + ζ(t), (3) 13

14 where, for instance, the drag coefficient α = 6πηa for a sphere of radius a moving in a fluid with viscosity η. Here, f represented a possible external force applied to the particle (e.g., by an electric or gravitational field). We established a fundamental connection between dissipation (specifically, the drag coefficient α here) and the fluctuating force ζ(t). For this, it was convenient to rewrite the equation of motion (6) in the absence of external forces. This allowed us to solve for the velocity in terms of an initial value v() and the fluctuating force ζ(t). We thus found a measurable quantity (specifically, v ) that depends directly on the random forces ζ. This dependence, however, is only through the correlation function K(t, t ) = ζ(t) ζ(t ), which has several general properties that identified above. Finally, we found that the drag coefficient α is fundamentally related to the random forces ζ, via Equation 31, which expresses the fact that the drag that a particle experiences in the fluid is due, at a microscopic level, to the randomly fluctuating forces that the particle feels at finite temperature. In other words, if we were to pull the particle through the fluid, the drag force experienced can be entirely expressed in terms of the forces ζ, which are presumably due to the random equilibrium fluctuations of the fluid! These random forces are at the origin of macroscopic dissipation..5 Spectral (Fourier) analysis of fluctuations When dealing with fluctuating quantities such as ζ and correlation functions such as K(t, t ) it is often most convenient to characterize them in terms of frequency. We have already noted that each specific realization of the random force ζ(t) (i.e., for each specific member of an ensemble of similar systems) is a very rapidly varying function, with a characteristic time scale of τ, which is very short compared with other time scales of interest, such as that of the motion of the particle (which also fluctuates). ζ(t) varies much more rapidly than does, for instance, v(t). Thus, ζ is expected to contain very high-frequency components (details) that are not apparent in the motion of our hypothetical particle moving around in the fluid. As we shall see, for all practical purposes, we can say that the frequencies characteristic of ζ are unbounded, or that it has components that vary arbitrarily rapidly in time. In order to make these notions more precise, let s look at the Fourier transforms of fluctuating quantities such as ζ(t). This can be done by using the orthogonality property of complex exponential functions (see Appendix C), 1 e iω(t t ) dω = δ (t t ), (33) π where δ(t t ) is the Dirac δ function, which is zero for t t, yet which integrates to unity: δ(t)dt = 1. 14

15 Consider some fluctuating quantity f(t) for a particular member of the ensemble. This could be, for instance, ζ itself, but we ll suppress any vector aspects. We can write In other words, where f(t) = = 1 π δ(t t )f(t )dt dt e iω(t t ) f(t )dω. f(t) = 1 f(ω)e iωt dω, π f(ω) = f(t )e iωt dt, (34) which define the Fourier transform and inverse Fourier transform. Here, f(ω) describes the frequency content of f(t). Assuming that f is real, then because the complex conjugate satisfies f (t) = f(t), (35) we also have that f (ω) = f( ω). (36) It is also worth noting that we have made some serious assumptions above concerning convergence of the various integrals. This will not pose a practical problem, but one might worry about this, since the function ζ is presumably nonzero and is fluctuating over the same range for arbitrarily large times. Thus, for instance, ζ(ω) might not be convergent. One could instead define the transforms for a modified function ζ Θ (t) that agrees with ζ(t) for t < Θ, and vanishes for t > Θ. One could then consider a limiting procedure in which Θ. We can do the same transform for the correlation function K(s) = ζ(t + s)ζ(t). We call the Fourier transform of this C(ω), and the two are related by and K(s) = 1 π C(ω) = C(ω)e iωs dω K(s )e iωs ds. (37) But, based on simple and general physical considerations, we showed before that K(t, t ) = ζ(t)ζ(t ) not only depends just on the time interval s = t t, but K(s) is also symmetric. It is also, of course, real. It is not hard to show that this also means that C(ω) is a real function and is symmetric in ω. We also argued before that K(s) is a function peaked at s = and rapidly decaying to zero within a time interval s of order τ, a very short molecular time scale. Thus, if ω is much smaller than 1/τ, then the argument of the exponential 15

16 in (37) is small wherever the K(s) is appreciable. This suggests that C(ω) is nearly independent of frequency ω for ω less than 1/τ, a very high frequency. In fact, such (molecular) frequencies are higher than any relevant frequency for the motion of our sphere in the fluid. So, we say that the noise spectrum is white. Just as we argued before that we can say that a constant. Note in particular that and C() = K(s) = Cδ(s), C(ω) = C, K(s )ds = 6kT α, K() = ζ() ζ() = 1 C(ω)dω π is of order C/τ. We can also relate the correlation functions K(s) and C(ω) to the Fourier transform of the noise ζ(ω). First of all, we note that (for most systems) we can replace the ensemble average for a single realization of the fluid or member of the ensemble in the definition of K by a time average over a wide enough time interval Θ < t < Θ: K(s) = ζ(t + s)ζ(t) In other words, = 1 Θ = 1 Θ = 1 Θ Θ Θ Θ Θ ζ(t + s)ζ(t) dω dω dt π e iω(t+s) ζ(ω) π e iω t ζ(ω ) dω π e iωs ζ(ω) ζ( ω). C(ω) = 1 Θ ζ(ω). Here, we have used the orthogonality of the complex exponential functions and the fact that ζ is a real function. It may seem strange that a measurable quantity C(ω), which we have seen is directly related to the drag coefficient, and thus the diffusion constant, seems to depend inversely on the time over which we observe the system Θ. However, as we have also noted, the Fourier transform ζ(ω) may not be well defined (e.g., finite). In fact, one should use the modified functions ζ Θ (t) described above. These have finite transforms that grow with Θ. But, the expression relating C(ω) to ζ Θ (ω) can be expected to become independent of Θ for large enough Θ (see Appendix D.1). 16

17 An important historical application of these ideas was to simple electrical circuits, which satisfy an equation similar to that of a particle in a dissipative medium: L d dt I(t) = V ext RI(t) + V (t), (38) where L is the inductance, I is the current, V ext represents a (constant or slowly varying) external applied field (emf), R is the resistance, and V is the fluctuating field related to dissipation. Taking the ensemble average of this equation yields the familiar law relating the voltage, current, and resistance: I = 1 R V ext. The resistance here is analogous to the drag coefficient α above, and a similar analysis to what was done above leads to R = 1 kt V (t + s)v (t) = C V V () kt. Here, the factor of rather than 6 is due to the fact that we are dealing with an inherently one-dimensional problem rather than a three-dimensional one. Furthermore, if the fluctuations of V are very fast (i.e., characterized by a short correlation time τ, here corresponding to the time between collisions of electrons in the wires, for instance), as for the fluid, then the spectrum of these fluctuations is very broad range C V V (ω) = V (t + s)v (t) e iωs ds = kt R for ω in a range of order 1/τ to 1/τ. This general relationship between the voltage fluctuations and the resistance in a circuit is known as Nyquist s theorem, and is yet another special case of the fluctuation-dissipation theorem. For a step-by-step derivation, see Appendix D.3. In Appendix D., we discuss another important noise type, namely shot noise. We have seen so far that the macroscopic drag coefficient, and thus the mobility and diffusion coefficient, of a particle in a simple liquid is related directly to the spectrum of microscopic fluctuations within the liquid. This also means that the physics governing the response of a system that is taken out of equilibrium (e.g., by dragging a sphere through a liquid) can be described entirely in terms if the fluctuations of the system about the equilibrium state. This is the fundamental observation behind the Fluctuation-Dissipation Theorem and the somewhat earlier Regression Hypothesis made by Lars Onsager. Before deriving these general relationships between fluctuations and response, however, let s examine one last aspect of our simple Langevin treatment of particle motion (fluctuations) within liquids. Namely, let us see how the mobility of the particle (e.g., when subjected to an external driving force) relates to the random fluctuations of the particle motion rather than the random fluid motion. We ll do this in just one dimension, and will thus neglect vector notation. 17

18 Consider the position of the particle, which we assume to start at t = from the origin. This satisfies Thus, r (t) = r(t) = Here, what we see is a correlation function v(t )dt. K vv (t, t ) v(t )v(t ) dt dt v(t )v(t ). (39) that is reminiscent of K we saw before. In fact, K vv satisfies all the properties that we previously identified for K = K ζζ, except that we can expect to find a longer correlation time of order τ. In particular, K vv should be a (symmetric) function only of the time interval between t and t. We can anticipate a result based on (39). Since we expect K vv (s) to decay toward zero for times long compared with the correlation time τ, if we are interested in times long compared with this time, then we expect that we can also approximate K vv (s) by a δ function, resulting in where r (t) = tcvv, C vv = K vv (s)ds is some constant. In other words, we expect to find diffusion for long times. We can, however, do better than approximating K vv (s) by a δ function. We know from before that Thus, v(t) = v()e t/τ + e t/τ 1 m v(t + s)v(t) = v() e (t+s)/τ + e (t+s)/τ m = v() e (t+s)/τ + e (t+s)/τ m where we have made the approximation made before. Continuing, we find that e t /τ ζ(t )dt. (4) (t+s) C ζζ ζ(t )ζ(t ) = C ζζ δ(t t ) v(t + s)v(t) = v() e (t+s)/τ + e (t+s)/τ m = v() e (t+s)/τ + e (t+s)/τ m 18 C ζζ C ζζ τ (t+s) e (t +t )/τ ζ(t )ζ(t ) dt dt min(t+s,t) e (t +t )/τ δ(t t )dt dt, e t /τ dt, { e t/τ 1 if s > e (t+s)/τ 1 otherwise

19 For times t long enough ( τ) so that all memory of the initial value of the velocity is lost, v(t + s)v(t) kt m e s /τ. (This is a one-dimensional result.) Indeed, for t τ, we find that where C vv = r (t) tc vv, v(t + s)v(t) ds = kt τ m, Which agrees with our prior result for the three-dimensional diffusion, once we account for the appropriate factors of 3: r (t) = 6Dt, where D = kt/α. Continuing in one dimension, however, we note the complementary expressions for the drag coefficient α and the mobility µ: and α = 1 kt µ = 1 α = 1 kt ζ(t + s)ζ(t) ds (41) v(t + s)v(t) ds. (4) For the electric circuit problem we introduced before, the analogous results for the resistance R and conductivity Σ are: and R = 1 kt Σ = 1 kt V (t + s)v (t) ds (43) I(t + s)i(t) ds, (44) where V is the (fluctuating) voltage and I is the corresponding current. The second of these relationships is usually written in terms of the current density j and the (local) conductivity σ. This is known as the Kubo formula σ = 1 kt j(t + s)j(t) ds. (45) In each of these cases, we have a pair of corresponding quantities: force and velocity, voltage and current,.... In each case, the pair of quantities is related in the sense that some generalized force (force or voltage) results in a response (drift velocity or current). In each case, the corresponding correlation functions, e.g. K ζζ and K vv, are related by an equation of the form ( ) ( ) ζ(t + s)ζ(t) ds v(t + s)v(t) ds = (kt ). (46) 19

20 .6 The Fluctuation-Dissipation Theorem and Linear Response We have seen in the examples above that one can, at least sometimes, relate quantities that have to do with systems out of equilibrium to the spectrum of thermal fluctuations about equilibrium. These principles are based on the Fluctuation-Dissipation Theorem, which was proven in general by Callen and Welton in 195, although special cases (as we have seen) of this were understood much earlier. The basic idea behind the fluctuation dissipation theorem was also captured in the 193s by the regression hypothesis of Onsager, who argued that the physics of macroscopic relaxation of a system back to equilibrium is governed by the same physics as the relaxation of spontaneous fluctuations about equilibrium. We have seen this, for instance in the relaxation of an initial value of v() of the velocity of a particle in a viscous liquid is described by the same time constant τ that appears in the correlation function K vv above. In order to derive the general relationship between, say the relaxation of some macroscopic observable and its equilibrium fluctuations, we need to look at the microscopic evolution of the system. This can either be done in the context of quantum mechanics, as is done in the book by Chaikin and Lubensky, or in the context of classical, Hamiltonian mechanics, as we shall do below. We describe the microscopic state of the system by a point in phase space described by a set of generalized momenta {p 1, p,...} and corresponding coordinates {q 1, q,...}. Although this is a very high-dimensional space (with of order 1 3 coordinates!), we ll denote is systematically by (p, q). The only essential result of Hamiltonian mechanics that we need to recall is the fact that, given some initial point in phase space (p(), q()) at time t =, the state of the system (p(t), q(t)) at time t in the future is completely determined by the initial condition and the Hamiltonian H (p(), q()). It is convenient for us to denote this evolution of the system by some time-evolution operator T t, defined by (p(t), q(t)) = T t (p(), q()). In classical statistical mechanics, the microstates of a system in equilibrium are distributed according to a probability distribution P (p, q) = 1 Q e βh(p,q), where Q = dp dq e βh(p,q) and β = 1/(kT ). The integral is over all of phase space. Again, we use an abbreviated notation, in which dp dq indicates an infinitesimal volume element in phase space. Consider what happens if we disturb the system, taking it out of equilibrium (or, more precisely, taking it to a new equilibrium state characterized by a new Hamiltonian H ). In fact, we shall look at the reverse problem, in which a perturbing field was applied to the system in the distant past, resulting in a new equilibrium at time t = characterized by the modified Hamiltonian H. At time t =, we turn off this field, and allow the system to relax back to equilibrium

21 (characterized by H). It is this relaxation of the system back to equilibrium that we want to examine. Let A(p, q) be some macroscopic observable whose relaxation we are interested in. We shall consider only linear response theory, in which we assume the perturbation is sufficiently weak that the Hamiltonian H can be taken to be H = H+ H, where H = fa and f is the perturbing field. (It is not hard to see that with this definition, f and A are conjugate thermodynamic variables, like pressure and volume, since f = F A, where F is the free energy.) The initial state of the system corresponds to distribution P (p, q) = 1 Q e βh (p,q), where Q = dp dq e βh (p,q). The initial (macro)state of the system is one in which A = 1 Q dp dq e βh (p,q) A(p, q). After the field is turned off, however, the system (i.e., each microstate (p, q)) evolves according to the Hamiltonian H(p, q) and the time-evolution operator T, rather than H (p, q) and T. In other words, at times t >, A(t) = 1 Q dp dq e βh (p,q) A (T t (p, q)) = dp dq e βh(p,q) (1 β H) A (T t (p, q)). dp dq e βh(p,q) (1 β H) Expanding, and keeping only terms through linear order gives A(t) dp dq e βh(p,q) A (T t (p, q)) = dp dq e βh(p,q) dp dq e βh(p,q) A(p, q)a (T t (p, q)) +βf dp dq e βh(p,q) dp dq e βh(p,q) A(p, q) dp dq e βf βh(p,q) A (T t (p, q)). dp dq e βh(p,q) dp dq e βh(p,q) This means that, to linear order, A(t) A(t) = βf ( A()A(t) A ), = βf δa()δa(t), (47) where δa(t) = A(t) A and refers to an ensemble average in the unperturbed system. We have been somewhat sloppy in the notation here. The first term in this equation refers to a evolution to time t of an ensemble of not only similarly prepared systems, but systems subject to the same perturbation. 1

22 It does not strictly refer to any ensemble average in the usual sense. Thus, it might be less confusing to write A(t) = βf δa()δa(t), (48) where this expresses the average deviation A from its equilibrium value. The average is to be understood to mean over the ensemble of systems with the same perturbation. This is one form of the Fluctuation-Dissipation theorem, expressing the macroscopic evolution of a system out of equilibrium to the spontaneous fluctuations of the system about equilibrium. This is usually expressed, however, in terms of the (linear) response of the system. This is implicit in the expression above. We have seen how the response A(t) above is linear in the force (or, more generally, the perturbation) f. This perturbation may, however, be time dependent. The most general form of a linear response to a time-dependent f is A(t) = dt χ(t, t )f(t ), where χ expresses the fact that the response at time t depends on the way the system was perturbed at other times. In principle, we should allow for arbitrarily influences from distant times. Thus, the integral above should be over all times t. It would be unphysical, however, if this influence extended to future times. How, for instance, can the system anticipate the perturbation that it will experience in the future? This simple, and physically reasonable assumption of causality can be expressed mathematically as χ(t, t ) = for all t > t. Note that χ in this sort of description is assumed to be a property of the (equilibrium) system, and not itself dependent on f. Thus, we also expect that χ should depend only on the time interval between t and t, so that { χ(t, t χ(t t ) = ) if t > t otherwise. As noted already, we expect that χ is a property of the (equilibrium) system, and not itself dependent on f. We should, therefore, be able to express it in terms if the dynamics of the equilibrium system. We have examined a particular case above, in which we perturbed the system for times t <, then removed the perturbation. This corresponds to { f if t < f(t) = otherwise. Here, we found that A(t) = βf δa()δa(t), (49) But, A(t) = dt χ(t, t )f

23 in this case, since f = for t >. By a change of variable of integration, this becomes A(t) = f t dt χ(t ). But, this only applies for t >. Thus, { β d χ(t) = dt δa()δa(t) if t > otherwise. This expresses the way in which a system responds to a perturbation (taking it out of equilibrium) in terms of the spontaneous fluctuations of the system about equilibrium. Last time, we saw how the response of the system, specifically in terms of some observable quantity A, could be described by a linear response function χ: A(t) = dt χ(t, t )f(t ), (5) where χ expresses the fact that the response at time t depends on the way the system was perturbed by some generalized force f at other times. Here, A(t) A(t) A, where A(t) represents the expected value of A at time t (this is actually an ensemble average quantity, but for a particular ensemble of similarly perturbed systems.) Based on the physical principle of causality, however, we expect the response only to depend on the history of the system and its perturbation. Furthermore, since the response function should be only an equilibrium property of the system itself (and not, for instance, dependent on f, at least if the perturbation is weak enough that the response is, indeed, linear). Thus, we expect that { χ(t, t χ(t t ) = ) if t > t otherwise. In deriving the above, we assumed a very particular form of the perturbed Hamiltonian: H = H fa. In general, however, we expect if the (generalized) force f is weak, then one can simply expand this Hamiltonian to linear order, with the result that H = H fb. That is, the linear coefficient of f need not be A, but may be some other variable B. When the system is subjected to such a perturbation, we expect to find a linear response of the variable A, as given by (5), where we shall denote the response function by χ AB (t t ), since it represents the response of A to a perturbation f that is conjugate to B. Here, f and B are conjugate in the 3

24 thermodynamic sense, in that F B = kt B log Q = kt 1 dp dq e β(h fb) Q B = f, where Q = dp dq e β(h fb) is the partition function and F = kt log Q is the free energy. This is just like the relationship between the conjugate variables volume V and pressure P : ( ) F = P. V T,N As before, if we consider a perturbing field f that is turned on in the distant past, and then switched off at time t =, we can write the evolution of (the ensemble average of) A as A(t) = 1 Q dp dq e βh (p,q) A (T t (p, q)) = dp dq e β(h(p,q) fb(p,q)) A (T t (p, q)). dp dq e β(h(p,q) fb(p,q)) An alternative way of deriving the linear response that we calculated before is to look at the derivative of this with respect to the perturbation f: = β Q dp dq e βh (p,q) B (p, q) A (T t (p, q)) A(t) f dp dq e βh (p,q) A (T t (p, q)) ( ) dp dq e βh (p,q) β dp dq e βh (p,q) B (p, q). Evaluating this derivative for f =, and noting that the point (p, q) in phase space represents the state of the system at time t =, we find that A(t) βf ( B()A(t) A(t) B() ) = βf B()δA(t). (51) Following the same analysis as before, we also find the response function χ AB (t) in terms of the fluctuating quantities δa and B: χ AB (t) = β d dt B()δA(t). (5) 4

25 Again, here, we should note that these thermodynamic averages refer to the unperturbed (i.e., equilibrium) system. Thus, they represent equilibrium fluctuations of the system, which exhibit some temporal correlations. The main example we have been looking at is the motion of a particle in a viscous fluid. Here, the perturbing field can be taken to be an external force f applied to the particle (say, by some magnetic or gravitational field). The response of the system is to develop a drift velocity v. But, the thermodynamically conjugate variable to f is the displacement x of the particle (in the direction of the force, which we take to be the x direction). Thus, H = H fx. So, the ensemble average drift velocity is v(t) = χ vx (t t )f(t )dt = f χ vx (t )dt, when, as before, a force f is applied in the distant past and switched off at time t =. But, in this case, we expect that v() = µf, where µ is the particle mobility. But, we know that µ = 1 kt v(t)v() dt = 1 kt t v(t)v() dt, where we have used the symmetry of K vv (t) = v(t)v(). This suggests that χ vx (t) = βk vv (t). This says that the dynamics associated with the response χ of a system to a non-equilibrium perturbation are governed by the same underlying principles as the fluctuations K vv (t) = v(t)v() about equilibrium. This is, indeed, the case, as Onsager suggested in his regression hypothesis. But, let s see if we can derive the relationship χ vx (t) = βk vv (t). We see from (5) that But, which cannot depend on t. Hence, χ vx (t) = β d dt x()v(t). (53) x()v(t) = x(t )v(t + t ), = d dt x(t )v(t + t ) = d dt x(t )v(t + t ) + x(t ) d dt v(t + t ) = v(t )v(t + t ) + x(t ) v(t + t ), and d dt x(t )v(t + t ) = x(t ) v(t + t ) = v(t )v(t + t ) = K vv (t). 5

26 This shows that χ vx (t) = βk vv (t) = 1 m e t/τ, for our particle in the fluid. We see that this represents what is sometimes referred to as a memory function. It shows how the response at one time depends on the history of disturbances, but with vanishing influence from distant past. That is, the memory fades over time. (As we all know too well!).7 The Damped Harmonic Oscillator The harmonic oscillator illustrates nicely many of the ideas we have been developing. It also turns out to be a very practical application of the Langevin analysis. There are many experimental situations of current interest where one has a harmonic potential for particles that would otherwise undergo brownian motion in a fluid. So, to our brownian particle above we add a harmonic, trapping potential Kx, which tends to localize the particle near x =. Our equation of motion becomes mẍ(t) + αẋ(t) + Kx(t) = f(t), where f is the force acting on the particle. We ll consider the case where this is some external driving force. Of course, there will also be our familiar random brownian forces. But, we shall treat ensemble quantities below. Fourier transforming the equation of motion leads to ( K mω iαω ) x ω = f ω, where we shall tend to employ a shorthand notation for transforms x ω of x(t), and similarly for the force. This can also be written in a way suggestive of linear response: 1 x ω = (K mω iαω) f ω, where 1 χ(ω) = K mω iαω is the Fourier transform of the response function defined by x(t) = dt χ(t t )f(t ). It is tempting at this point to anticipate the frequency-dependent version of our expression which would read χ xx (t) = β d dt x(t)x() = β d dt K xx(t), (54) χ(ω) = iωβ x ω x ω. (incorrect) 6

15.3 The Langevin theory of the Brownian motion

15.3 The Langevin theory of the Brownian motion 5.3 The Langevin theory of the Brownian motion 593 Now, by virtue of the distribution (2), we obtain r(t) =; r 2 (t) = n(r,t)4πr 4 dr = 6Dt t, (22) N in complete agreement with our earlier results, namely

More information

CHAPTER V. Brownian motion. V.1 Langevin dynamics

CHAPTER V. Brownian motion. V.1 Langevin dynamics CHAPTER V Brownian motion In this chapter, we study the very general paradigm provided by Brownian motion. Originally, this motion is that a heavy particle, called Brownian particle, immersed in a fluid

More information

Brownian Motion and Langevin Equations

Brownian Motion and Langevin Equations 1 Brownian Motion and Langevin Equations 1.1 Langevin Equation and the Fluctuation- Dissipation Theorem The theory of Brownian motion is perhaps the simplest approximate way to treat the dynamics of nonequilibrium

More information

This is a Gaussian probability centered around m = 0 (the most probable and mean position is the origin) and the mean square displacement m 2 = n,or

This is a Gaussian probability centered around m = 0 (the most probable and mean position is the origin) and the mean square displacement m 2 = n,or Physics 7b: Statistical Mechanics Brownian Motion Brownian motion is the motion of a particle due to the buffeting by the molecules in a gas or liquid. The particle must be small enough that the effects

More information

4. The Green Kubo Relations

4. The Green Kubo Relations 4. The Green Kubo Relations 4.1 The Langevin Equation In 1828 the botanist Robert Brown observed the motion of pollen grains suspended in a fluid. Although the system was allowed to come to equilibrium,

More information

Physics 562: Statistical Mechanics Spring 2003, James P. Sethna Homework 5, due Wednesday, April 2 Latest revision: April 4, 2003, 8:53 am

Physics 562: Statistical Mechanics Spring 2003, James P. Sethna Homework 5, due Wednesday, April 2 Latest revision: April 4, 2003, 8:53 am Physics 562: Statistical Mechanics Spring 2003, James P. Sethna Homework 5, due Wednesday, April 2 Latest revision: April 4, 2003, 8:53 am Reading David Chandler, Introduction to Modern Statistical Mechanics,

More information

Non-equilibrium phenomena and fluctuation relations

Non-equilibrium phenomena and fluctuation relations Non-equilibrium phenomena and fluctuation relations Lamberto Rondoni Politecnico di Torino Beijing 16 March 2012 http://www.rarenoise.lnl.infn.it/ Outline 1 Background: Local Thermodyamic Equilibrium 2

More information

F(t) equilibrium under H 0

F(t) equilibrium under H 0 Physics 17b: Statistical Mechanics Linear Response Theory Useful references are Callen and Greene [1], and Chandler [], chapter 16. Task To calculate the change in a measurement B t) due to the application

More information

Brownian motion and the Central Limit Theorem

Brownian motion and the Central Limit Theorem Brownian motion and the Central Limit Theorem Amir Bar January 4, 3 Based on Shang-Keng Ma, Statistical Mechanics, sections.,.7 and the course s notes section 6. Introduction In this tutorial we shall

More information

t = no of steps of length s

t = no of steps of length s s t = no of steps of length s Figure : Schematic of the path of a diffusing molecule, for example, one in a gas or a liquid. The particle is moving in steps of length s. For a molecule in a liquid the

More information

Linear Response and Onsager Reciprocal Relations

Linear Response and Onsager Reciprocal Relations Linear Response and Onsager Reciprocal Relations Amir Bar January 1, 013 Based on Kittel, Elementary statistical physics, chapters 33-34; Kubo,Toda and Hashitsume, Statistical Physics II, chapter 1; and

More information

G : Statistical Mechanics

G : Statistical Mechanics G25.2651: Statistical Mechanics Notes for Lecture 15 Consider Hamilton s equations in the form I. CLASSICAL LINEAR RESPONSE THEORY q i = H p i ṗ i = H q i We noted early in the course that an ensemble

More information

Topics covered so far:

Topics covered so far: Topics covered so far: Chap 1: The kinetic theory of gases P, T, and the Ideal Gas Law Chap 2: The principles of statistical mechanics 2.1, The Boltzmann law (spatial distribution) 2.2, The distribution

More information

2.3 Damping, phases and all that

2.3 Damping, phases and all that 2.3. DAMPING, PHASES AND ALL THAT 107 2.3 Damping, phases and all that If we imagine taking our idealized mass on a spring and dunking it in water or, more dramatically, in molasses), then there will be

More information

Brownian Motion and The Atomic Theory

Brownian Motion and The Atomic Theory Brownian Motion and The Atomic Theory Albert Einstein Annus Mirabilis Centenary Lecture Simeon Hellerman Institute for Advanced Study, 5/20/2005 Founders Day 1 1. What phenomenon did Einstein explain?

More information

Onsager theory: overview

Onsager theory: overview Onsager theory: overview Pearu Peterson December 18, 2006 1 Introduction Our aim is to study matter that consists of large number of molecules. A complete mechanical description of such a system is practically

More information

08. Brownian Motion. University of Rhode Island. Gerhard Müller University of Rhode Island,

08. Brownian Motion. University of Rhode Island. Gerhard Müller University of Rhode Island, University of Rhode Island DigitalCommons@URI Nonequilibrium Statistical Physics Physics Course Materials 1-19-215 8. Brownian Motion Gerhard Müller University of Rhode Island, gmuller@uri.edu Follow this

More information

The dynamics of small particles whose size is roughly 1 µmt or. smaller, in a fluid at room temperature, is extremely erratic, and is

The dynamics of small particles whose size is roughly 1 µmt or. smaller, in a fluid at room temperature, is extremely erratic, and is 1 I. BROWNIAN MOTION The dynamics of small particles whose size is roughly 1 µmt or smaller, in a fluid at room temperature, is extremely erratic, and is called Brownian motion. The velocity of such particles

More information

Major Concepts Lecture #11 Rigoberto Hernandez. TST & Transport 1

Major Concepts Lecture #11 Rigoberto Hernandez. TST & Transport 1 Major Concepts Onsager s Regression Hypothesis Relaxation of a perturbation Regression of fluctuations Fluctuation-Dissipation Theorem Proof of FDT & relation to Onsager s Regression Hypothesis Response

More information

LINEAR RESPONSE THEORY

LINEAR RESPONSE THEORY MIT Department of Chemistry 5.74, Spring 5: Introductory Quantum Mechanics II Instructor: Professor Andrei Tokmakoff p. 8 LINEAR RESPONSE THEORY We have statistically described the time-dependent behavior

More information

Noise and Brownian motion. So far, the equations we have been studying in the context of nonequilibrium statistical

Noise and Brownian motion. So far, the equations we have been studying in the context of nonequilibrium statistical Phys504 Statistical Physics Fall 2006 Lecture #20 Professor Anthony J. Leggett Department of Physics, UIUC Noise and Brownian motion So far, the equations we have been studying in the context of nonequilibrium

More information

Collective Effects. Equilibrium and Nonequilibrium Physics

Collective Effects. Equilibrium and Nonequilibrium Physics Collective Effects in Equilibrium and Nonequilibrium Physics: Lecture 5, April 14, 2006 1 Collective Effects in Equilibrium and Nonequilibrium Physics Website: http://cncs.bnu.edu.cn/mccross/course/ Caltech

More information

Damped harmonic motion

Damped harmonic motion Damped harmonic motion March 3, 016 Harmonic motion is studied in the presence of a damping force proportional to the velocity. The complex method is introduced, and the different cases of under-damping,

More information

In-class exercises Day 1

In-class exercises Day 1 Physics 4488/6562: Statistical Mechanics http://www.physics.cornell.edu/sethna/teaching/562/ Material for Week 11 Exercises due Mon Apr 16 Last correction at April 16, 2018, 11:19 am c 2018, James Sethna,

More information

Time-Dependent Statistical Mechanics 5. The classical atomic fluid, classical mechanics, and classical equilibrium statistical mechanics

Time-Dependent Statistical Mechanics 5. The classical atomic fluid, classical mechanics, and classical equilibrium statistical mechanics Time-Dependent Statistical Mechanics 5. The classical atomic fluid, classical mechanics, and classical equilibrium statistical mechanics c Hans C. Andersen October 1, 2009 While we know that in principle

More information

Collective Effects. Equilibrium and Nonequilibrium Physics

Collective Effects. Equilibrium and Nonequilibrium Physics 1 Collective Effects in Equilibrium and Nonequilibrium Physics: Lecture 5, April 14, 2006 1 Collective Effects in Equilibrium and Nonequilibrium Physics Website: http://cncs.bnu.edu.cn/mccross/course/

More information

2 Equations of Motion

2 Equations of Motion 2 Equations of Motion system. In this section, we will derive the six full equations of motion in a non-rotating, Cartesian coordinate 2.1 Six equations of motion (non-rotating, Cartesian coordinates)

More information

UNIVERSITY OF OSLO FACULTY OF MATHEMATICS AND NATURAL SCIENCES

UNIVERSITY OF OSLO FACULTY OF MATHEMATICS AND NATURAL SCIENCES UNIVERSITY OF OSLO FCULTY OF MTHEMTICS ND NTURL SCIENCES Exam in: FYS430, Statistical Mechanics Day of exam: Jun.6. 203 Problem :. The relative fluctuations in an extensive quantity, like the energy, depends

More information

2. Molecules in Motion

2. Molecules in Motion 2. Molecules in Motion Kinetic Theory of Gases (microscopic viewpoint) assumptions (1) particles of mass m and diameter d; ceaseless random motion (2) dilute gas: d λ, λ = mean free path = average distance

More information

III. Kinetic Theory of Gases

III. Kinetic Theory of Gases III. Kinetic Theory of Gases III.A General Definitions Kinetic theory studies the macroscopic properties of large numbers of particles, starting from their (classical) equations of motion. Thermodynamics

More information

Drude theory & linear response

Drude theory & linear response DRAFT: run through L A TEX on 9 May 16 at 13:51 Drude theory & linear response 1 Static conductivity According to classical mechanics, the motion of a free electron in a constant E field obeys the Newton

More information

1 Brownian motion and the Langevin equation

1 Brownian motion and the Langevin equation Figure 1: The robust appearance of Robert Brown (1773 1858) 1 Brownian otion and the Langevin equation In 1827, while exaining pollen grains and the spores of osses suspended in water under a icroscope,

More information

BROWNIAN DYNAMICS SIMULATIONS WITH HYDRODYNAMICS. Abstract

BROWNIAN DYNAMICS SIMULATIONS WITH HYDRODYNAMICS. Abstract BROWNIAN DYNAMICS SIMULATIONS WITH HYDRODYNAMICS Juan J. Cerdà 1 1 Institut für Computerphysik, Pfaffenwaldring 27, Universität Stuttgart, 70569 Stuttgart, Germany. (Dated: July 21, 2009) Abstract - 1

More information

Noise, AFMs, and Nanomechanical Biosensors

Noise, AFMs, and Nanomechanical Biosensors Noise, AFMs, and Nanomechanical Biosensors: Lancaster University, November, 2005 1 Noise, AFMs, and Nanomechanical Biosensors with Mark Paul (Virginia Tech), and the Caltech BioNEMS Collaboration Support:

More information

Brownian Motion: Fokker-Planck Equation

Brownian Motion: Fokker-Planck Equation Chapter 7 Brownian Motion: Fokker-Planck Equation The Fokker-Planck equation is the equation governing the time evolution of the probability density of the Brownian particla. It is a second order differential

More information

kg meter ii) Note the dimensions of ρ τ are kg 2 velocity 2 meter = 1 sec 2 We will interpret this velocity in upcoming slides.

kg meter ii) Note the dimensions of ρ τ are kg 2 velocity 2 meter = 1 sec 2 We will interpret this velocity in upcoming slides. II. Generalizing the 1-dimensional wave equation First generalize the notation. i) "q" has meant transverse deflection of the string. Replace q Ψ, where Ψ may indicate other properties of the medium that

More information

Chapter 2. Dielectric Theories

Chapter 2. Dielectric Theories Chapter Dielectric Theories . Dielectric Theories 1.1. Introduction Measurements of dielectric properties of materials is very important because it provide vital information regarding the material characteristics,

More information

A path integral approach to the Langevin equation

A path integral approach to the Langevin equation A path integral approach to the Langevin equation - Ashok Das Reference: A path integral approach to the Langevin equation, A. Das, S. Panda and J. R. L. Santos, arxiv:1411.0256 (to be published in Int.

More information

Lecture 11: Long-wavelength expansion in the Neel state Energetic terms

Lecture 11: Long-wavelength expansion in the Neel state Energetic terms Lecture 11: Long-wavelength expansion in the Neel state Energetic terms In the last class we derived the low energy effective Hamiltonian for a Mott insulator. This derivation is an example of the kind

More information

MD Thermodynamics. Lecture 12 3/26/18. Harvard SEAS AP 275 Atomistic Modeling of Materials Boris Kozinsky

MD Thermodynamics. Lecture 12 3/26/18. Harvard SEAS AP 275 Atomistic Modeling of Materials Boris Kozinsky MD Thermodynamics Lecture 1 3/6/18 1 Molecular dynamics The force depends on positions only (not velocities) Total energy is conserved (micro canonical evolution) Newton s equations of motion (second order

More information

Continuum Mechanics Fundamentals

Continuum Mechanics Fundamentals Continuum Mechanics Fundamentals James R. Rice, notes for ES 220, 12 November 2009; corrections 9 December 2009 Conserved Quantities Let a conseved quantity have amount F per unit volume. Examples are

More information

J10M.1 - Rod on a Rail (M93M.2)

J10M.1 - Rod on a Rail (M93M.2) Part I - Mechanics J10M.1 - Rod on a Rail (M93M.2) J10M.1 - Rod on a Rail (M93M.2) s α l θ g z x A uniform rod of length l and mass m moves in the x-z plane. One end of the rod is suspended from a straight

More information

Figure 1: Doing work on a block by pushing it across the floor.

Figure 1: Doing work on a block by pushing it across the floor. Work Let s imagine I have a block which I m pushing across the floor, shown in Figure 1. If I m moving the block at constant velocity, then I know that I have to apply a force to compensate the effects

More information

in order to insure that the Liouville equation for f(?; t) is still valid. These equations of motion will give rise to a distribution function f(?; t)

in order to insure that the Liouville equation for f(?; t) is still valid. These equations of motion will give rise to a distribution function f(?; t) G25.2651: Statistical Mechanics Notes for Lecture 21 Consider Hamilton's equations in the form I. CLASSICAL LINEAR RESPONSE THEORY _q i = @H @p i _p i =? @H @q i We noted early in the course that an ensemble

More information

Biomolecular hydrodynamics

Biomolecular hydrodynamics Biomolecular hydrodynamics Chem 341, Fall, 2014 1 Frictional coefficients Consider a particle moving with velocity v under the influence of some external force F (say a graviational or electrostatic external

More information

Chapter 6: The Rouse Model. The Bead (friction factor) and Spring (Gaussian entropy) Molecular Model:

Chapter 6: The Rouse Model. The Bead (friction factor) and Spring (Gaussian entropy) Molecular Model: G. R. Strobl, Chapter 6 "The Physics of Polymers, 2'nd Ed." Springer, NY, (1997). R. B. Bird, R. C. Armstrong, O. Hassager, "Dynamics of Polymeric Liquids", Vol. 2, John Wiley and Sons (1977). M. Doi,

More information

5.74 Introductory Quantum Mechanics II

5.74 Introductory Quantum Mechanics II MIT OpenCourseWare ttp://ocw.mit.edu 5.74 Introductory Quantum Mecanics II Spring 9 For information about citing tese materials or our Terms of Use, visit: ttp://ocw.mit.edu/terms. Andrei Tokmakoff, MIT

More information

Non equilibrium thermodynamics: foundations, scope, and extension to the meso scale. Miguel Rubi

Non equilibrium thermodynamics: foundations, scope, and extension to the meso scale. Miguel Rubi Non equilibrium thermodynamics: foundations, scope, and extension to the meso scale Miguel Rubi References S.R. de Groot and P. Mazur, Non equilibrium Thermodynamics, Dover, New York, 1984 J.M. Vilar and

More information

Polymer dynamics. Course M6 Lecture 5 26/1/2004 (JAE) 5.1 Introduction. Diffusion of polymers in melts and dilute solution.

Polymer dynamics. Course M6 Lecture 5 26/1/2004 (JAE) 5.1 Introduction. Diffusion of polymers in melts and dilute solution. Course M6 Lecture 5 6//004 Polymer dynamics Diffusion of polymers in melts and dilute solution Dr James Elliott 5. Introduction So far, we have considered the static configurations and morphologies of

More information

Equations of linear stellar oscillations

Equations of linear stellar oscillations Chapter 4 Equations of linear stellar oscillations In the present chapter the equations governing small oscillations around a spherical equilibrium state are derived. The general equations were presented

More information

Chapter 9: Statistical Mechanics

Chapter 9: Statistical Mechanics Chapter 9: Statistical Mechanics Chapter 9: Statistical Mechanics...111 9.1 Introduction...111 9.2 Statistical Mechanics...113 9.2.1 The Hamiltonian...113 9.2.2 Phase Space...114 9.2.3 Trajectories and

More information

Physics 4022 Notes on Density Matrices

Physics 4022 Notes on Density Matrices Physics 40 Notes on Density Matrices Definition: For a system in a definite normalized state ψ > the density matrix ρ is ρ = ψ >< ψ 1) From Eq 1 it is obvious that in the basis defined by ψ > and other

More information

Fundamental equations of relativistic fluid dynamics

Fundamental equations of relativistic fluid dynamics CHAPTER VI Fundamental equations of relativistic fluid dynamics When the energy density becomes large as may happen for instance in compact astrophysical objects, in the early Universe, or in high-energy

More information

Short Wire Antennas: A Simplified Approach Part I: Scaling Arguments. Dan Dobkin version 1.0 July 8, 2005

Short Wire Antennas: A Simplified Approach Part I: Scaling Arguments. Dan Dobkin version 1.0 July 8, 2005 Short Wire Antennas: A Simplified Approach Part I: Scaling Arguments Dan Dobkin version 1.0 July 8, 2005 0. Introduction: How does a wire dipole antenna work? How do we find the resistance and the reactance?

More information

Langevin Methods. Burkhard Dünweg Max Planck Institute for Polymer Research Ackermannweg 10 D Mainz Germany

Langevin Methods. Burkhard Dünweg Max Planck Institute for Polymer Research Ackermannweg 10 D Mainz Germany Langevin Methods Burkhard Dünweg Max Planck Institute for Polymer Research Ackermannweg 1 D 55128 Mainz Germany Motivation Original idea: Fast and slow degrees of freedom Example: Brownian motion Replace

More information

Rubber elasticity. Marc R. Roussel Department of Chemistry and Biochemistry University of Lethbridge. February 21, 2009

Rubber elasticity. Marc R. Roussel Department of Chemistry and Biochemistry University of Lethbridge. February 21, 2009 Rubber elasticity Marc R. Roussel Department of Chemistry and Biochemistry University of Lethbridge February 21, 2009 A rubber is a material that can undergo large deformations e.g. stretching to five

More information

Consider a parallel LCR circuit maintained at a temperature T >> ( k b

Consider a parallel LCR circuit maintained at a temperature T >> ( k b UW Physics PhD. Qualifying Exam Spring 8, problem Consider a parallel LCR circuit maintained at a temperature T >> ( k b LC ). Use Maxwell-Boltzmann statistics to calculate the mean-square flux threading

More information

Linear Theory of Evolution to an Unstable State

Linear Theory of Evolution to an Unstable State Chapter 2 Linear Theory of Evolution to an Unstable State c 2012 by William Klein, Harvey Gould, and Jan Tobochnik 1 October 2012 2.1 Introduction The simple theory of nucleation that we introduced in

More information

Joel A. Shapiro January 20, 2011

Joel A. Shapiro January 20, 2011 Joel A. shapiro@physics.rutgers.edu January 20, 2011 Course Information Instructor: Joel Serin 325 5-5500 X 3886, shapiro@physics Book: Jackson: Classical Electrodynamics (3rd Ed.) Web home page: www.physics.rutgers.edu/grad/504

More information

Lecture 1: Pragmatic Introduction to Stochastic Differential Equations

Lecture 1: Pragmatic Introduction to Stochastic Differential Equations Lecture 1: Pragmatic Introduction to Stochastic Differential Equations Simo Särkkä Aalto University, Finland (visiting at Oxford University, UK) November 13, 2013 Simo Särkkä (Aalto) Lecture 1: Pragmatic

More information

5.74 Introductory Quantum Mechanics II

5.74 Introductory Quantum Mechanics II MIT OpenCourseWare http://ocw.mit.edu 5.74 Introductory Quantum Mechanics II Spring 9 For information about citing these materials or our Terms of Use, visit: http://ocw.mit.edu/terms. Andrei Tokmakoff,

More information

Symmetry of the Dielectric Tensor

Symmetry of the Dielectric Tensor Symmetry of the Dielectric Tensor Curtis R. Menyuk June 11, 2010 In this note, I derive the symmetry of the dielectric tensor in two ways. The derivations are taken from Landau and Lifshitz s Statistical

More information

MACROSCOPIC VARIABLES, THERMAL EQUILIBRIUM. Contents AND BOLTZMANN ENTROPY. 1 Macroscopic Variables 3. 2 Local quantities and Hydrodynamics fields 4

MACROSCOPIC VARIABLES, THERMAL EQUILIBRIUM. Contents AND BOLTZMANN ENTROPY. 1 Macroscopic Variables 3. 2 Local quantities and Hydrodynamics fields 4 MACROSCOPIC VARIABLES, THERMAL EQUILIBRIUM AND BOLTZMANN ENTROPY Contents 1 Macroscopic Variables 3 2 Local quantities and Hydrodynamics fields 4 3 Coarse-graining 6 4 Thermal equilibrium 9 5 Two systems

More information

G : Statistical Mechanics Notes for Lecture 3 I. MICROCANONICAL ENSEMBLE: CONDITIONS FOR THERMAL EQUILIBRIUM Consider bringing two systems into

G : Statistical Mechanics Notes for Lecture 3 I. MICROCANONICAL ENSEMBLE: CONDITIONS FOR THERMAL EQUILIBRIUM Consider bringing two systems into G25.2651: Statistical Mechanics Notes for Lecture 3 I. MICROCANONICAL ENSEMBLE: CONDITIONS FOR THERMAL EQUILIBRIUM Consider bringing two systems into thermal contact. By thermal contact, we mean that the

More information

Linearization of Differential Equation Models

Linearization of Differential Equation Models Linearization of Differential Equation Models 1 Motivation We cannot solve most nonlinear models, so we often instead try to get an overall feel for the way the model behaves: we sometimes talk about looking

More information

( ) f (k) = FT (R(x)) = R(k)

( ) f (k) = FT (R(x)) = R(k) Solving ODEs using Fourier Transforms The formulas for derivatives are particularly useful because they reduce ODEs to algebraic expressions. Consider the following ODE d 2 dx + p d 2 dx + q f (x) = R(x)

More information

Collective behavior, from particles to fields

Collective behavior, from particles to fields 978-0-51-87341-3 - Statistical Physics of Fields 1 Collective behavior, from particles to fields 1.1 Introduction One of the most successful aspects of physics in the twentieth century was revealing the

More information

Time-Dependent Statistical Mechanics 1. Introduction

Time-Dependent Statistical Mechanics 1. Introduction Time-Dependent Statistical Mechanics 1. Introduction c Hans C. Andersen Announcements September 24, 2009 Lecture 1 9/22/09 1 Topics of concern in the course We shall be concerned with the time dependent

More information

Before we consider two canonical turbulent flows we need a general description of turbulence.

Before we consider two canonical turbulent flows we need a general description of turbulence. Chapter 2 Canonical Turbulent Flows Before we consider two canonical turbulent flows we need a general description of turbulence. 2.1 A Brief Introduction to Turbulence One way of looking at turbulent

More information

Contents. 1 Introduction 1

Contents. 1 Introduction 1 Abstract In the first part of this thesis we review the concept of stochastic Langevin equations. We state a simple example and explain its physical meaning in terms of Brownian motion. We then calculate

More information

The formulas for derivatives are particularly useful because they reduce ODEs to algebraic expressions. Consider the following ODE d 2 dx + p d

The formulas for derivatives are particularly useful because they reduce ODEs to algebraic expressions. Consider the following ODE d 2 dx + p d Solving ODEs using Fourier Transforms The formulas for derivatives are particularly useful because they reduce ODEs to algebraic expressions. Consider the following ODE d 2 dx + p d 2 dx + q f (x) R(x)

More information

The Euler Equation of Gas-Dynamics

The Euler Equation of Gas-Dynamics The Euler Equation of Gas-Dynamics A. Mignone October 24, 217 In this lecture we study some properties of the Euler equations of gasdynamics, + (u) = ( ) u + u u + p = a p + u p + γp u = where, p and u

More information

Physics 172H Modern Mechanics

Physics 172H Modern Mechanics Physics 172H Modern Mechanics Instructor: Dr. Mark Haugan Office: PHYS 282 haugan@purdue.edu TAs: Alex Kryzwda John Lorenz akryzwda@purdue.edu jdlorenz@purdue.edu Lecture 22: Matter & Interactions, Ch.

More information

Modelling across different time and length scales

Modelling across different time and length scales 18/1/007 Course MP5 Lecture 1 18/1/007 Modelling across different time and length scales An introduction to multiscale modelling and mesoscale methods Dr James Elliott 0.1 Introduction 6 lectures by JAE,

More information

2nd-Order Linear Equations

2nd-Order Linear Equations 4 2nd-Order Linear Equations 4.1 Linear Independence of Functions In linear algebra the notion of linear independence arises frequently in the context of vector spaces. If V is a vector space over the

More information

Statistical Mechanics

Statistical Mechanics 42 My God, He Plays Dice! Statistical Mechanics Statistical Mechanics 43 Statistical Mechanics Statistical mechanics and thermodynamics are nineteenthcentury classical physics, but they contain the seeds

More information

3. General properties of phase transitions and the Landau theory

3. General properties of phase transitions and the Landau theory 3. General properties of phase transitions and the Landau theory In this Section we review the general properties and the terminology used to characterise phase transitions, which you will have already

More information

C C C C 2 C 2 C 2 C + u + v + (w + w P ) = D t x y z X. (1a) y 2 + D Z. z 2

C C C C 2 C 2 C 2 C + u + v + (w + w P ) = D t x y z X. (1a) y 2 + D Z. z 2 This chapter provides an introduction to the transport of particles that are either more dense (e.g. mineral sediment) or less dense (e.g. bubbles) than the fluid. A method of estimating the settling velocity

More information

The Principle of Least Action

The Principle of Least Action The Principle of Least Action Anders Svensson Abstract In this paper, the principle of least action in classical mechanics is studied. The term is used in several different contexts, mainly for Hamilton

More information

CHAPTER 9. Microscopic Approach: from Boltzmann to Navier-Stokes. In the previous chapter we derived the closed Boltzmann equation:

CHAPTER 9. Microscopic Approach: from Boltzmann to Navier-Stokes. In the previous chapter we derived the closed Boltzmann equation: CHAPTER 9 Microscopic Approach: from Boltzmann to Navier-Stokes In the previous chapter we derived the closed Boltzmann equation: df dt = f +{f,h} = I[f] where I[f] is the collision integral, and we have

More information

Once Upon A Time, There Was A Certain Ludwig

Once Upon A Time, There Was A Certain Ludwig Once Upon A Time, There Was A Certain Ludwig Statistical Mechanics: Ensembles, Distributions, Entropy and Thermostatting Srinivas Mushnoori Chemical & Biochemical Engineering Rutgers, The State University

More information

Statistical Mechanics in a Nutshell

Statistical Mechanics in a Nutshell Chapter 2 Statistical Mechanics in a Nutshell Adapted from: Understanding Molecular Simulation Daan Frenkel and Berend Smit Academic Press (2001) pp. 9-22 11 2.1 Introduction In this course, we will treat

More information

Langevin Dynamics of a Single Particle

Langevin Dynamics of a Single Particle Overview Langevin Dynamics of a Single Particle We consider a spherical particle of radius r immersed in a viscous fluid and suppose that the dynamics of the particle depend on two forces: A drag force

More information

LANGEVIN THEORY OF BROWNIAN MOTION. Contents. 1 Langevin theory. 1 Langevin theory 1. 2 The Ornstein-Uhlenbeck process 8

LANGEVIN THEORY OF BROWNIAN MOTION. Contents. 1 Langevin theory. 1 Langevin theory 1. 2 The Ornstein-Uhlenbeck process 8 Contents LANGEVIN THEORY OF BROWNIAN MOTION 1 Langevin theory 1 2 The Ornstein-Uhlenbeck process 8 1 Langevin theory Einstein (as well as Smoluchowski) was well aware that the theory of Brownian motion

More information

Lecture 3: From Random Walks to Continuum Diffusion

Lecture 3: From Random Walks to Continuum Diffusion Lecture 3: From Random Walks to Continuum Diffusion Martin Z. Bazant Department of Mathematics, MIT February 3, 6 Overview In the previous lecture (by Prof. Yip), we discussed how individual particles

More information

Resonance and response

Resonance and response Chapter 2 Resonance and response Last updated September 20, 2008 In this section of the course we begin with a very simple system a mass hanging from a spring and see how some remarkable ideas emerge.

More information

Time dependent perturbation theory 1 D. E. Soper 2 University of Oregon 11 May 2012

Time dependent perturbation theory 1 D. E. Soper 2 University of Oregon 11 May 2012 Time dependent perturbation theory D. E. Soper University of Oregon May 0 offer here some background for Chapter 5 of J. J. Sakurai, Modern Quantum Mechanics. The problem Let the hamiltonian for a system

More information

CHAPTER 7 SEVERAL FORMS OF THE EQUATIONS OF MOTION

CHAPTER 7 SEVERAL FORMS OF THE EQUATIONS OF MOTION CHAPTER 7 SEVERAL FORMS OF THE EQUATIONS OF MOTION 7.1 THE NAVIER-STOKES EQUATIONS Under the assumption of a Newtonian stress-rate-of-strain constitutive equation and a linear, thermally conductive medium,

More information

10. Zwanzig-Mori Formalism

10. Zwanzig-Mori Formalism University of Rhode Island DigitalCommons@URI Nonequilibrium Statistical Physics Physics Course Materials 205 0. Zwanzig-Mori Formalism Gerhard Müller University of Rhode Island, gmuller@uri.edu Creative

More information

Summer Lecture Notes Thermodynamics: Fundamental Relation, Parameters, and Maxwell Relations

Summer Lecture Notes Thermodynamics: Fundamental Relation, Parameters, and Maxwell Relations Summer Lecture Notes Thermodynamics: Fundamental Relation, Parameters, and Maxwell Relations Andrew Forrester August 4, 2006 1 The Fundamental (Difference or Differential) Relation of Thermodynamics 1

More information

The Truth about diffusion (in liquids)

The Truth about diffusion (in liquids) The Truth about diffusion (in liquids) Aleksandar Donev Courant Institute, New York University & Eric Vanden-Eijnden, Courant In honor of Berni Julian Alder LLNL, August 20th 2015 A. Donev (CIMS) Diffusion

More information

Waves in a Shock Tube

Waves in a Shock Tube Waves in a Shock Tube Ivan Christov c February 5, 005 Abstract. This paper discusses linear-wave solutions and simple-wave solutions to the Navier Stokes equations for an inviscid and compressible fluid

More information

7 The Navier-Stokes Equations

7 The Navier-Stokes Equations 18.354/12.27 Spring 214 7 The Navier-Stokes Equations In the previous section, we have seen how one can deduce the general structure of hydrodynamic equations from purely macroscopic considerations and

More information

Linear-response theory and the fluctuation-dissipation theorem: An executive summary

Linear-response theory and the fluctuation-dissipation theorem: An executive summary Notes prepared for the 27 Summer School at Søminestationen, Holbæk, July 1-8 Linear-response theory and the fluctuation-dissipation theorem: An executive summary Jeppe C. Dyre DNRF centre Glass and Time,

More information

CERN Accelerator School. Intermediate Accelerator Physics Course Chios, Greece, September Low Emittance Rings

CERN Accelerator School. Intermediate Accelerator Physics Course Chios, Greece, September Low Emittance Rings CERN Accelerator School Intermediate Accelerator Physics Course Chios, Greece, September 2011 Low Emittance Rings Part 1: Beam Dynamics with Synchrotron Radiation Andy Wolski The Cockcroft Institute, and

More information

Cosmology and particle physics

Cosmology and particle physics Cosmology and particle physics Lecture notes Timm Wrase Lecture 9 Inflation - part I Having discussed the thermal history of our universe and in particular its evolution at times larger than 10 14 seconds

More information

PHYS 352. Noise. Noise. fluctuations in voltage (or current) about zero average voltage = 0 average V 2 is not zero

PHYS 352. Noise. Noise. fluctuations in voltage (or current) about zero average voltage = 0 average V 2 is not zero PHYS 352 Noise Noise fluctuations in voltage (or current) about zero average voltage = 0 average V 2 is not zero so, we talk about rms voltage for noise V=0 1 Sources of Intrinsic Noise sometimes noise

More information

G : Statistical Mechanics

G : Statistical Mechanics G25.2651: Statistical Mechanics Notes for Lecture 1 Defining statistical mechanics: Statistical Mechanics provies the connection between microscopic motion of individual atoms of matter and macroscopically

More information

Introduction Statistical Thermodynamics. Monday, January 6, 14

Introduction Statistical Thermodynamics. Monday, January 6, 14 Introduction Statistical Thermodynamics 1 Molecular Simulations Molecular dynamics: solve equations of motion Monte Carlo: importance sampling r 1 r 2 r n MD MC r 1 r 2 2 r n 2 3 3 4 4 Questions How can

More information

Lecture 1: The Equilibrium Green Function Method

Lecture 1: The Equilibrium Green Function Method Lecture 1: The Equilibrium Green Function Method Mark Jarrell April 27, 2011 Contents 1 Why Green functions? 2 2 Different types of Green functions 4 2.1 Retarded, advanced, time ordered and Matsubara

More information