M.Phys. C6 Theory Option Lecture Notes For 2020-2021
M.Phys. C6 Theory Option Lecture Notes For 2020-2021
C6 Theory Option
Lecture Notes for 2020-2021
Lectures given by Sid Parameswaran in Michaelmas 2020
Lecture Notes written by F.H.L. Essler with contributions by Sid Parameswaran
The Rudolf Peierls Centre for Theoretical Physics
Oxford University, Oxford OX1 3PU, UK
November 4, 2020
These notes cover Part V of the lectures (Random Systems and Stochastic Processes), which is only part
of M.Phys. C6 Theory Option.
Please report errors and typos to [email protected].
2015
c F.H.L. Essler.
Contents
1 Random Variables 3
1.1 Some Definitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.2 Discrete-time random walk . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.3 The central limit theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
3 Markov Processes 6
3.1 Examples of Markov Processes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
3.2 Markov Chains . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
4 Brownian Motion 9
4.1 Langevin Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
4.2 Fokker-Planck Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
4.3 Diffusion Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
1
• Basic ideas of probability, kinetic theory, etc.
M. Kardar, Statistical Physics of Particles, Cambridge.
• Stochastic Processes
N.G. van Kampen, Stochastic Processes in Physics and Chemistry, North-Holland/Elsevier.
2
Part V
Random Systems and Stochastic Processes
In nature there are many phenomena, in which some quantity varies in a random way. An example is
Browninan motion, which refers to the motion of a small particle suspended in a fluid. The motion, observed
under a microscope, looks random. It is hopeless to try to compute the position in detail, but certain average
features obey simple laws. Averaging over a suitable time interval is difficult and one therefore replaces time
averaging of a single irregularly varying function of time by averaging over an ensemble of functions. The
latter must of course be chosen in such a way that the two results agree
1 Random Variables
A random variable is an object X defined by
1
{xj } = {1, 2, 3, 4, 5, 6} , PX (xj ) = . (3)
6
The average value of X is the first moment hXi, while the variance is σ 2 = hX 2 i − hXi2 .
3
The Fourier transform of PX (x) is called the characteristic function
∞
(ik)n
Z X
φX (k) = dx PX (x)eikx = hX n i ≡ heikX i.
n!
n=0
(5)
The last equality shows that the characteristic function is the generating function for the moments
dn
hX n i = (−i)n n φX (k). (6)
dk k=0
The cumulants of PX (x) are defined as
dn
Cn (X) = (−i)n
n
ln φX (k) .
dk k=0
(7)
Clearly we have
hY i = 0 , hY 2 i = N hXj2 i = N , (11)
where we have used that the steps are mutually independent. To obtain pN (n) we employ the characteristic
function
φXj (k) = PXj (1)eik1 + PXj (−1)eik(−1) = cos(k). (12)
The characteristic function of the random variable Y is
PN N
Y N
φY (k) = heikY i = heik j=1 Xj
i= heikXj i = φXj (k)
j=1
N
1 ik −ik
N 1 X N ik(N −2r)
= e +e = N e . (13)
2N 2 r
r=0
4
On the other hand we have by definition of the characteristic function
X N
X
ikn
φY (k) = PY (n)e = pN (n)eikn . (14)
n n=−N
5
2 Stochastic Processes: Definitions
A function YX (t) of time t and a random variable X is called a stochastic process (SP). Examples are
the position x(t) or the velocity v(t) of a particle undergoing Brownian motion. A SP is characterized by
probability densities
These are
• normalized Z
dy1 . . . dyn Pn (y1 , t1 ; . . . ; yn , tn ) = 1 ; (25)
• reducible Z
dyn Pn (y1 , t1 ; . . . ; yn , tn ) = Pn−1 (y1 , t1 ; . . . ; yn−1 , tn−1 ). (26)
This is normalized Z
dy2 P1|1 (y2 , t2 |y1 , t1 ) = 1 , (30)
3 Markov Processes
Perhaps the most important stochastic processes are so-called Markov processes (MP). Their defining prop-
erty is that t1 < t2 < . . . < tn
6
This means that at time tn−1 one can predict the state of the system at time tn on the basis of present
information, i.e. yn−1 , only! The history of how the system arrived at yn−1 at time tn−1 is irrelevant.
A Markov process is completely determined by the two functions P1 (y1 , t1 ) and P1|1 (y2 , t2 |y1 , t1 ), and
this makes Markov processes tractable. Any P1 (y1 , t1 ) and P1|1 (y2 , t2 |y1 , t1 ) define a MP, provided that they
fulfil the following two consistency conditions
• Chapman-Kolmogorov equation
Z
P1|1 (y3 , t3 |y1 , t1 ) = dy2 P1|1 (y3 , t3 |y2 , t2 )P1|1 (y2 , t2 |y1 , t1 ) , t3 > t2 > t1 .
(34)
• Evolution equation
Z
P1 (y2 , t2 ) = dy1 P1|1 (y2 , t2 |y1 , t1 )P1 (y1 , t1 ) , t2 > t1 .
(35)
A MP is called stationary if P1 (y, t) is time-independent and P1|1 (y2 , t2 |y1 , t1 ) depends only on the time
difference t2 − t1 (and y1,2 ).
N
X
Tjk = 1,
j=1
(37)
P1|1 (1, t + 1; 1, t) = 1 − q ,
P1|1 (2, t + 1; 1, t) = q ,
P1|1 (1, t + 1; 2, t) = r ,
P1|1 (2, t + 1; 2, t) = 1 − r . (38)
7
Introducing a two-component vector
P1 (1, t)
p~(t) = , (39)
P1 (2, t)
the evolution equation for the process can be expressed as a vector equation
1−q r
p~(t + 1) = p~(t).
q 1−r
| {z }
T
(40)
T is a square matrix with non-negative entries, that is in general not symmetric. The matrix elements Tij
are the rates for transitions from state j to state i. It is useful to rewrite the equation in components
X
pn (t + 1) = Tnj pj (t). (41)
j
Then X X X
pn (t + 1) − pn (t) = Tnj pj (t) − Tjn pn (t) = Tnj pj (t) − Tjn pn (t). (42)
j j j
| {z }
1
Now consider our time interval to be small instead of 1. Then (42) turns into a differential equation, called
a Master equation
dpn (t) X
= Tnj pj (t) − Tjn pn (t).
dt
j
(43)
This has a nice physical interpretation as a “loss/gain” equation for probabilities: the first term on the
right-hand side is the rate of transitions from state j to state n times the probability of j being realized, i.e.
the total gain of probability for state n. The second term on the right-hand side is the rate of transitions
from state n to state j times the probability of n being realized, i.e. the total loss of probability for state n.
Let us now return to the discrete form (40), which can be iterated to give
While T is generally not symmetric, it is nevertheless often diagonalizable. Then there exist left and right
eigenvectors such that
In our example
1 r/q
λ1 = 1 , hL1 | = (1, 1) , |R1 i = . (48)
1 + r/q 1
q 1 −1
λ2 = 1 − q − r , hL2 | = (− , 1) , |R2 i = . (49)
r 1 + r/q 1
8
So for large t we have
t+1 1 r r
T ≈ |R1 ihL1 | = , (50)
r+q q q
and hence
1 r r
p~(∞) = p~(0). (51)
r+q q q
4 Brownian Motion
We now want to think of Brownian motion as a Markov process. Let v1 , v2 , . . . be the velocities of the
particle at different time steps. Then vk+1 depends only on vk , but not on v1 , . . . , vk−1 .
dv(t)
= −γv(t) + η(t).
dt (52)
Here the first term on the right hand side is damping term linear in v, while the second term represents the
remaining random force with zero average hη(t)i = 0. This is often referred to as “noise”. For simplicity we
will assume collisions to be instantaneous, so that forces at different times are uncorrelated
hη(t)η(t0 )i = Γδ(t − t0 ). (53)
Given our assumptions about the noise, we can calculate noise-averaged quantities quite easily. We have
d 0 γt0 dv 0 0
0
v(t )e = 0
+ γv eγt = η(t0 )eγt , (54)
dt dt
where in the last step we used the Langevin equation (52). Integrating both sides of this equation between
t = 0 and t0 = t, we obtain
Z t
−γt 0
v(t) = v(0)e + dt0 η(t0 )e−γ(t−t ) .
0
(55)
Averaging this over the noise, we find the average velocity
Z t
0
hv(t)i = v(0)e−γt + dt0 hη(t0 )i e−γ(t−t ) = v(0)e−γt .
0 | {z }
=0
(56)
Similarly we have
Z t Z t
2 −γt 0 0 −γ(t−t0 ) −γt 00 00 −γ(t−t00 )
hv (t)i = h v(0)e + dt η(t )e v(0)e + dt η(t )e i
0 0
Z t
0 00
= v 2 (0)e−2γt + e−2γt dt0 dt00 hη(t0 )η(t00 )i eγ(t +t ) (57)
0 | {z }
Γδ(t0 −t00 )
Γ
hv 2 (t)i = v 2 (0)e−2γt + (1 − e−2γt ).
2γ
(58)
9
The displacement of the particle is
Z t Z t Z t0
0 0 v(0) −γt 0 0 00
x(t) − x(0) = dt v(t ) = (1 − e ) + dt dt00 η(t00 )e−γ(t −t ) . (59)
0 γ 0 0
v(0)
hx(t)i = (1 − e−γt ).
γ
(60)
2 Γ Γ Γ
h x(t) − hx(t)i i = 2 t − 3 (1 − e−γt ) − 3 (1 − e−γt )2 .
γ γ 2γ
(63)
Γ kB T
= .
2γ m
(67)
10
It is convenient to consider the integral
Z
Ω = dv [P1 (v, t + ∆t) − P1 (v, t)] h(v), (70)
where h(v) is is test function (infinitely many time differentiable, h(v) and all of its derivatives going to zero
at infinity etc). On the one hand, we have to linear order in ∆t
Z Z
∂P1 (v, t)
dv [P1 (v, t + ∆t) − P1 (v, t)] h(v) = dv ∆t h(v). (71)
∂t
Relabelling
R the integration variable from v to u in the second term, and using that normalization condition
dv P1|1 (v, t + ∆t|u, t) = 1, we obtain
Z Z
Ω = du P1 (u, t) dv P1|1 (v, t + ∆t|u, t) [h(v) − h(u)] . (73)
Integrating the n’th term in the sum n times by parts, and using the nice properties of the function h(u),
then leads to the following expression
∞
∂ n
Z X
Ω= du h(u) − P1 (u, t)D(n) (u). (75)
∂u
n=1
Using that (71) and (75) have to be equal for any test function h(u), we conclude that
∞
∂ n
∂P1 (v, t) X
∆t = − P1 (v, t)D(n) (v).
∂t ∂v
n=1
(76)
This starts looking like our desired differential equation. What remains is to determine the quantities D(n) (v)
(w − v)n zn
Z Z
(n)
D (v) = dw P1|1 (w, t + ∆t|v, t) = dz P1|1 (v + z, t + ∆t|v, t)
n! n!
1
= h[v(t + ∆t) − v(t)]n i. (77)
n!
We see that D(n) (v) are related to the moments of the velocity difference distribution! We can use the
Langevin equation to determine them, and the result is
11
Substituting these into (76) and then taking the limit ∆t → 0, we arrive at the Fokker-Planck equation
∂ ∂ Γ ∂2
P1 (v, t) = γ vP1 (v, t) + P1 (v, t).
∂t ∂v 2 ∂v 2 (79)
This is a second order linear PDE for P1 (v, t) and can be solved by standard methods. For initial conditions
P1 (v, 0) = δ(v − v0 ) we find
2 !
1 v − v̄(t)
P1 (v, t) = p exp − ,
2πσ 2 (t) 2σ 2 (t)
(80)
where
Γ
σ 2 (t) = (1 − e−2γt ) , v̄(t) = v0 e−γt . (81)
2γ
In the limit t → ∞ this turns into the Maxwell distribution
r r 2
γ − γv2 m − mv
P1 (v, t) = e Γ = e 2kB T . (82)
πΓ 2πkB T
Let us see how to derive (78), starting from the Langevin equation
Z t
0
v(t) = v(0)e−γt + dt0 e−γ(t−t ) η(t0 ). (83)
0
where
v0 (t) = −γv(0)e−γt ,
0 0
K(t, t0 ) = −γΘ(t − t0 )Θ(t0 )e−γ(t−t ) ∆t + Θ(t0 − t)Θ(t + ∆t − t0 )e−γ(t−t ) . (87)
Let us now assume that η(t) is Gaussian distributed. Then the probability distribution for the noise
is the functional
1 2
R
P [η(t)] = e− 2Γ dt η (t) . (88)
12
Averages are then given by the path integral
Z
1
R 0
η 2 (t0 )
hη(t1 ) . . . η(tn )i = Dη(t) η(t1 ) . . . η(tn ) e− 2Γ dt . (89)
hη(t)i = 0 ,
hη(t)η(t0 )i = Γδ(t − t0 ) ,
hη(t1 )η(t2 )η(t3 )i = hη(t1 )η(t2 )ihη(t3 )i + hη(t3 )η(t1 )ihη(t2 )i + hη(t2 )η(t3 )ihη(t1 )i = 0 ,
hη(t1 )η(t2 )η(t3 )η(t4 )i = hη(t1 )η(t2 )ihη(t3 )η(t4 )i + . . . (90)
The probability distribution for ∆v(t) can be obtained from the generating function
Z
1
R 0 2 0 0 0
Z(λ) = heλ∆v(t) i = eλv0 (t)∆t Dη(t) e− 2Γ dt [η (t )−2λΓK(t,t )η(t )] . (91)
Changing variables to
η̃(t0 ) = η(t0 ) − λΓK(t, t0 ), (92)
the generating function becomes
λ2 Γ
dt0 K 2 (t,t0 )
R
Z(λ) = eλv0 (t)∆t+ 2 . (93)
Γ −γ|t−t0 |
hv(t)i = 0 , hv(t)v(t0 )i = e . (96)
2γ
Now let us imagine that we observe the Brownian particle only at sufficiently long time intervals t, t0 γ −1 ,
and describe only these coarse grained positions. Then we may replace
Γ −γ|t−t0 | Γ
hv(t)v(t0 )i = e → 2 δ(t − t0 ). (97)
2γ γ
13
Γ −γ|t−t0 |
This is because 2γ e is substantially different from zero only if |t − t0 | < γ −1 , and
Z ∞
Γ −γ|t−t0 | Γ
dt e = 2. (98)
−∞ 2γ γ
dx(t)
= v(t), (99)
dt
then turns into a special case (of no damping) of the Langevin equation we solved for the velocity. We
therefore can use our previous results to conclude that
∂ Γ ∂2
P1 (x, t) = 2 2 P1 (x, t).
∂t 2γ ∂x
(100)
2 !
1 x − x0
P1 (x, t) = p exp − .
2πD|t − t0 | 4D|t − t0 |
(102)
Observe that the diffusion equation is a special case of the Fokker-Planck equation in the absence of a
viscous damping term (the fact that there is a γ in the expression should not worry you – the equation for
x has no damping term, which is the important thing.)
where in the second step, we have chosen to work in the position basis. Evidently, there is a formal resem-
blance between equations (103) and (105): namely, P1|1 (y, t|y0 , t0 ) plays a role analogous to the propagator
U (t; t0 ) in quantum mechanics. Given the state of the system at a given time t0 , it ‘propagates’ this forward
in time to time t > t0 . Of course, (105) describes probability amplitudes rather than probabilities, and
14
therein lies a key difference between quantum and classical physics. However, we can exploit the formal
resemblance between the classical evolution of probabilities in a Markov process and the quantum evolution
of amplitudes under unitary time evolution to recast the former in path integral form. We now carry out
this procedure, specializing to the case of Brownian motion.
Recall that the fundamental idea in the path integral approach to QM was to write the time evolution
operator as a path integral by inserting a complete set of states after each infinitesimal time step. The
analogous idea for the Markov process is to repeatedly use the Chapman-Kolmogorov equation to split up
the time evolution into infinitesimal chunks. Let us split the time interval from t0 to t into N equal time
slices, defining = t−t N , tn = t0 + n, so that tN = t. Repeatedly applying (34), we have
0
Z
P1|1 (y, t|y0 , t0 ) = dy1 dy2 . . . dyN −1 P1|1 (y, t|yN −1 , tN −1 ) . . . P1|1 (y2 , t2 |y1 , t1 )P1|1 (y1 , t1 |y0 , t0 ) (106)
Z N
Y −1
= dy1 dy2 . . . dyN −1 P1|1 (yn+1 , tn+1 |yn , tn ). (107)
n=0
Just as for the QM case, we have broken up the evolution into infinitesimal evolutions. The reason is that it’s
frequently easier to compute the infinitesimal time evolution. Let’s consider the case of Brownian motion,
where we have from (36) that
−yn )2
1 (y
− 2(tn+1 1 − (y(tn+1 )−y(tn ))2
P1|1 (yn+1 , tn+1 |yn , tn ) = p e n+1 −tn ) =√ e 2 , (108)
2π(tn+1 − tn ) 2π
where in the second step we have used the fact that in our case tn − tn−1 = by construction, and defined
yn = y(tn ). Substituting this into (107) and rewriting, we have
N −1
1 − (y(tn+1 )−y(tn ))2
Z Y
P1|1 (y, t|y0 , t0 ) = dy1 dy2 . . . dyN −1 √ e 2
n=0
2π
N Z PN −1 (y(tn+1 )−y(tn ))2
1
= √ dy1 dy2 . . . dyN −1 e− n=0 2 . (109)
2π
(y(tn+1 )−y(tn ))
We now let N → ∞, which defines a trajectory or path y(t). We recognise once again that '
ẏ(tn ), and hence that
N −1
ẏ(t0 )2
Z t
X (y(tn+1 ) − y(tn ))2
' dt0 . (110)
2 0 2
n=0
Thus we find that we have expressed the propagator for Brownian motion as a path integral,
Z R t 0 ẏ(t0 )2
P1|1 (y, t|y0 , t0 ) = N Dy(t0 )e− 0 dt 2 . (111)
15
becomes
−(n− < n >)2
1
PN (n) = √ exp
2πσ 2 2σ 2
where σ 2 = N pq. Check that the same result follows from the central limit theorem.
(ii) Consider a random walk in one dimension, for which the probability of taking a step of length x → x + dx is
1 γ
f (x)dx = dx.
π x + γ2
2
Find the probability distribution for the total displacement after N steps. Does it satisfy the central limit theorem?
Should it? What are the cumulants of this distribution?
Question 24. This question is about a continuous random walk, also known as a Wiener process.
Show that for −∞ < y < ∞ and t2 > t1 the Chapman-Kolmogarov equation is satisfied for
(y2 − y1 )2
1
P1|1 (y2 , t2 | y1 , t1 ) = p exp − . (114)
2π(t2 − t1 ) 2(t2 − t1 )
−y 2
1
P1 (y, t) = √ exp . (115)
2πt 2t
∂P ∂2P
=D 2 (116)
∂t ∂y
Question 25. A particle suspended in a fluid undergoes Brownian motion in one dimension with position
x(t) and velocity v(t). This motion is modelled by the Langevin equation
dv
= −γv + η(t),
dt
where η(t) is a Gaussian random variable characterised completely by the averages hη(t)i = 0 and hη(t1 )η(t2 )i =
Γδ(t1 − t2 ). Discuss the physical origin of each of the terms in the Langevin equation.
What is meant by the term Markov process? Illustrate your answer by discussing which of the following are
Markov processes: (a) v(t) alone; (b) x(t) alone; (c) v(t) and x(t) together.
Show that for t > 0
Z t Z t1
v(0) −γt
x(t) = (1 − e ) + dt1 dt2 e−γ(t1 −t2 ) η(t2 )
γ 0 0
16
is a solution of the Langevin equation with initial condition x(0) = 0. Calculate the average hx(t) v(t)i and
discuss its limiting behaviour at short and long times.
Question 26. The time evolution of a stochastic system is represented by a master equation of the form
dpn (t) X
= Wnm pm (t) .
dt m
Explain briefly the meaning of this equation and discuss the assumptions on which it is based. What general
conditions should the matrix elements Wnm satisfy?
A molecule lies between two atomic-scale contacts and conducts charge between them. A simple model of
this situation is illustrated below. The model has three states: the molecule may be uncharged, or may carry a
single charge at either site A or site B but not both. Charges hop between these sites, and between the sites and
the contacts, at the rates indicated in the figure. (For example, a charge at site A has probability f2 per unit
time of hopping to site B.)
f1 f2 f3
A B
Write down a master equation for this model. For the system in equilibrium, calculate the occupation probabilities
of the three states, and show that the average number of charges flowing through the molecule per unit time is
f1 f2 f3
.
f1 f2 + f1 f3 + f2 f3
Consider the case f1 = f2 = f3 ≡ f . The molecule is uncharged at time t = 0. Show that the probability
p(t) for it to be uncharged at a later time t is
√ !
1 2 3 3
p(t) = + exp − f t cos ft .
3 3 2 2
17