New Course Notes
New Course Notes
Contents
8 Discrete-time dynamics 27
18 Numerical methods 75
18.1 Euler’s method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
18.2 Runge-Kutta method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
18.3 von Neumann rejection method for random number generation . . . . . . . . . 75
18.4 Tau-leaping . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
18.5 First-reaction and next-reaction methods . . . . . . . . . . . . . . . . . . . . . 75
Part I
Applied Mathematics:
Dynamics, Data, and Information
1 Introduction
Applied mathematics is in an interesting era: “Data” and “information” are the buzz words of
the day. The analytical skills needed for dealing with data, statistics and computer science,
are farther and farther dissociated from the central theme of traditional applied mathematics:
Understanding the solutions of ordinary and partial differential equations (ODEs & PDEs) with
analytical and numerical methods. In the Part I of this course, however, we shall learn a new
outlook and logic, and methods coming with it, of applied mathematics that encompasses and
significantly generalizes the traditional applied mathematics: The new logic is based on the
mathematical theory of probability; the application of which is an organic integration of the
modern theory of dynamical systems and classical statistics in terms of empirical frequencies.
The latter has been more or less abandoned by the field of statistics proper in the pursuing of
Bayesian statistics.1
The organic integration is due to the fact that the theory of dynamical systems is based on
the deep mathematical notion of measure, and the theory of probability is based on exactly
the same mathematics: probability is a normalized measure according to its founder, A. N.
Kolmogorov. Quote a statement from A. Ya. Khinchin, a disciple of the former, in his master
piece2 :
This theory [of dynamical systems] from a formal point of view could be considered
as a group of special problems of the theory of measures, namely such problems
as most often deal with the establishment of a metrically negligible smallness of
certain sets. It is suffices to remember that the majority of propositions of the
theory of functions of a real variable concerned with the notions of convergence
“in measure”, “almost everywhere” etc., finds an adequate expression in the
terminology of the theory of probability.
Perhaps, the deepest distinction between the applied mathematics tradition and the statistics
is that the former is based on “pursuing truth without the durden of being true”. The “being
true” part is supposely the job of physics who provides us with the law(s) of nature in terms of
ODEs or PDEs, and functional relations needed between quantities called constitutive relations.
Statisticians, on the other hand, have to work with numbers from measurements in the real
world, where none of the following exist: infinity, infinitesimal, real number most of which are
irrational, derivative from 00 , and of course measure and probability.
1
S. B. McGrayne (2012) The Theory That World Not Die: How Bayes’ rule cracked the enigma code, hunted
down Russian submarines, and emerged triumphant from two centuries of controversy, Yale UniversityPress, New
Haven.
2
A. I. Khinchin (1949) Mathematical Foundations of Statistical Mechanics, Dover, New York.
What is the justification of adding two measurements X1 and X2 ? Why should not αX1 +βX2 ?
By choosing α = β, there is in fact a hidden assumption of certain “identity”, or “equality”,
between the originators of the X1 and X2 .
The importance of a “reference set” and its choice should be based on the very question one
asks: How is the “prediction” going to be used?
Statistical inferences are based only in part upon the observations. An equally
important base is formed by prior assumptions about the underlying situation.
Even in the simplest cases, there are explicit or implicit assumptions about randomness
and independence, about distributional models, perhaps prior distributions for
some unknown parameters, and so on.
To understand this issue, let us work out a simple problem: N toss of a coin with k heads and
(N − k) tails. What is the conditional probability for the coin head?
In this problem, let Ω = {0, 1}. This means we have completely neglected the possibility
of a coin standing on its side. Let us denote the head as X = 1 and tail as X = 0. Assuming
all the tosses can be modeled as independent and identically distributed (i.i.d.) events, with
X1 , X2 , · · · , XN , what is
P Xi = 1 X1 + X2 + · · · + XN = k =? (2)
3
P. J. Huber and E. M. Ronchetti (2009) Robust Statistics, 2nd ed., John Wiley & Sons, New York, p. 1.
where 1 ≤ i ≤ N . That is, if the probability of the coin results in a head in a toss is P(X =
1) = p, then
P(X = 1, X + · · · + X = k − 1)
1 2 N
P X1 = 1 X1 + X2 · · · + XN = k =
P(X1 + X2 + · · · + XN = k)
(N − 1)!
p pk−1 (1 − p)N −k
(k − 1)!(N − k)! k
= = .
N! N
pk (1 − p)N −k
k!(N − k)!
Since all the toss are identical, the result will be the same for other Xi . How to explain this
mathematical result? Let us consider the following three scenarios.
Scenario I: John has a coin in hand. He throws it 10 times, and observes 1 head and 9
tails. Should John conclude that (i) this coin was very badly made, or (ii) that he simply had a
“unlucky streak”?
Scenario II: Jean has ten such coins. She tosses each one, and nine of them land on a
table with tails up and only the fifth one with head up. Looking at the coins on the table, she
concludes that what ever the p, exists or not, has no relevance any more.
Scenario III: Jessie is told by Jean that on the table there is exactly one coin with head
up but not which one. With that incomplete information in hand Jessie concludes based on
symmetry that each coin on the table has 9 to 1 odds being tail up.
John is uncertain between “frequentist” vs. “Bayesian”, Jean is a follower of Kolmogorov,
and Jessie is a statistical physicist, with whom we applied mathematic workers are in the same
camp. As stated by E. T. Jaynes in the preface of his masterpiece:4 “[F]or general problems
of scientific inference, almost all of which arise out of imcomplete information rather than
‘randomness’.”
Logistic map
un+1 = f (xt ) = run (1 − un ), 0 < r ≤ 4. (3)
Its unique, invariant measure with density when r = 4.5
4
E. T. Jaynes (2003) Probability Theory: The Logic of Science, Cambridge University Press.
5
L.-S. Young (2002) What are SRB measures, and which dynamical systems have them? Journal of Statistical
Physics, vol. 108, 733–754.
To understand this simple mathematical model which arises in insect population dynamics6 ,
we pay particular attention to the effect of the only parameter r in the model on the behavior
of the dymamics. The r is assumed between 1 and 4.
Because r ∈ [1, 4], we note that if un ∈ [0, 1] then un+1 ∈ [0, 1]. By dynamics, we mean
starting from a point u0 ∈ [0, 1], the sequence of number in “discrete time” n:
u0 , u1 , u2 , · · · , u100 , · · · , u1000 , · · · ,
as n → ∞. Eq. 3 is called a discrete time nonlinear dynamical system.
In the limit of n → ∞, the method of cobwebbing tells us that as long as r < 1, the un → 0
as n → ∞ for all u0 ∈ [0, 1]. When 1 < r < 3, however, the un → r−1 r
for all u0 ∈ [0, 1]
∗ ∗ ∗
except u0 = 0. A u ∈ [0, 1] that satisfies f (u ) = u is called a fixed point, or a steady state.
Clearly there are two types of steady state, stable and unstable.
√
Interestingly, when 3 < r < 1 + 6 ≈ 3.44949, there is one stable steady state r−1 and 0 is
√ r
always unstable. Then when r > 1 + 6, both of them are unstable, and a new type of “long-
time behavior” emerge: There is a pair of u∗1 and u∗2 such that f (u∗1 ) = u∗2 and f (u∗2 ) = u∗1 ,
and this goes on forever. This behavior is called a “2-cycle”. There is a very complex story
following this, and eventually when r = 4, we have unstable 0 and 43 , as well as unstable
2-cycles √ √
5 − 5 5 + 5
u∗1 = and u∗2 = ,
2 2
and infinitely many other such points in the interval [0, 1]. The infinitely many points made any
two initial values, say u0 and u′0 , no matter how close, diverge quickly as n → ∞. This is the
origin of “random behavior”; it is known as deterministic chaos.
This randomness can now be mathematically described as having a probability distribution
with probability density functionfor the value of u’s:
1
ρ(u) = p . (4)
π u(1 − u)
The sequence of numbers {un }, with r = 4, is equivalent to to a random number exccept those
unstable points. For any scientific perspective, those numbers are not relevant.
The function in Eq. (4) is known as the invariant probability density function. It can be shown
that many complex dynamical systems behave this way, with a unique invariant probability
density.
One can thus empirically obtain such an invariant distribution from continuous sampling
of the system ad infinitum. When the data is sufficiently “big”, one can determine the ρ(u)
function pretty well. This is the scientific significane of the mathematical concept of chaos.
6
R. M. May, (1976) Simple mathematical models with very complicated dynamics . Nature, vol. 261, 459–
467.
(i) Protein molecule, single protein and single polypeptide with random sequences;
(ii) An “eternal” cell, phenotypes, differentiation, etc.;
(iii) The amazing story of stem cell;
(iii) The cellular landscape;
(iv) Darwinian evolutionary process — recurrent vs. non-stationary;
(v) The chemical view of the world: Atoms as fundamental particles;
(vi) The transmutation between man and butterfly.
The mathematical theory of large deviations theory provides an axiomatic definition of “entropy”,
the elusive concept originated in thermal physics.
3 Energy:
An alternative representation of idealized statistics
3.1 Discrete finite state space Ω
3.2 Gedankenexperiment and holographic information
3.2.1 Empirical mean value of repeated measurements as a project
Part II
Mathematical Models, Theories and
Analyses in Biology
6 Re-Introduction
6.1 The philosophy of mathematical modeling in biology
Probability is the logic of science7 ; statistics is a tool for making discoveries from data; and
dynamics is a language for narrating exact science. In modern mathematics, probability does
not necessarily mean stochastic; it is about measure-theoretical representation of a subject.
Newtonian world view is a clockwise machine; complex many-body systems always have
“dissipation”. So where is the biological diversity coming from? You should keep this very
important question in mind all through this course.
rates, e.g., the amount of water going into and coming out the bathtub per unit time. In the
banking language: W (t) is the amount of money in the account, I(t) is the rate of deposite,
and O(t) is the rate of expanse.
rate of population increase = birth rate − death rate + immigration rate. (5)
If we use x(t) to denote the population at time t, then the above equation becomes
dx
= x b(x) − d(x) + i(x), (6)
dt
in which b and d are the per capita birth and death rates, respectively. Note that one of the
most important aspects of birth and death is that if x = 0, then there will be no possibility of
further birth or death. Without immigration, an extinct population will remain extinct. The
immigration term i(x), however, has a very different feature: It needs not to be zero when
x = 0.
Consider a population with many subpopulations ⃗x = (x1 , x2 , · · · , xn ), all xi ≥ 0. In the
absence of immigration, if we denote ri (⃗x) = bi (⃗x) − d(⃗x), then
dxi
= xi ri (⃗x), (7)
dt
and the per capita growth rate for the entire population, which is also the mean per capita
growth rate,
n n
X dxi X
xi ri (⃗x)
i=1
dt i=1
r= n = n , xi ≥ 0. (8)
X X
xi xi
i=1 i=1
Then,
"P
n Pn 2 # Pn x x r ∂ri
dr(⃗x) xi ri2 xi ri i,j=1 i j j ∂xj
= Pi=1
n − Pi=1
n + Pn (9)
dt i=1 xi i=1 xi i=1 xi
“As every individual, therefore, endeavours as much as he can both to employ his
capital in the support of domestic industry, and so to direct that industry that its
produce may be of the greatest value; every individual necessarily labours to render
the annual revenue of the society as great as he can. He generally, indeed, neither
intends to promote the public interest, nor knows how much he is promoting it.
By preferring the support of domestic to that of foreign industry, he intends only
his own security; and by directing that industry in such a manner as its produce
may be of the greatest value, he intends only his own gain, and he is in this, as in
many other eases, led by an invisible hand to promote an end which was no part
of his intention. Nor is it always the worse for the society that it was no part of
it. By pursuing his own interest he frequently promotes that of the society more
effectually than when he really intends to promote it. I have never known much
good done by those who affected to trade for the public good. It is an affectation,
indeed, not very common among merchants, and very few words need be employed
in dissuading them from it.”
Pn 2
i=1 xi ri − r
Pn Pn dri (⃗
x)
d xi ri i=1 xi dt
Pi=1
n − P n = Pn . (11)
dt i=1 x i i=1 x i i=1 xi
This equation can be phrased as “the change in the per capita growth rate of an entire population
is never less than the average change in per capita growth rate of the subpopulations”.9 Eq. 9
also shows that dr/dt could be negative if the last term on the right-hand-side is large and
negative. Therefore, it is interesting to investigate under what circumstances it is positive or
negative.
First, we note that if all ri are constant, independent of ⃗x, then this last term is zero since
∂ri /∂xj = 0.
Second, if ri is a linear function of ⃗x: ri (⃗x) = nk=1 wik xk . Furthermore, one can always
P
8
Edwards, A.W.F (1994) The fundamental theorem of natural selection. Biol. Rev. 69, 443–474.
9
Price, G.R. (1972) Fisher’s ‘fundamental theorem’ made clear. Ann. Hum. Genet. Lond. 36, 129–140.
S A
decompose a matrix into a symmetric and an anti-symmetric parts: wij = wij + wij . Then
Pn
∂ri Pn
x x r
i,j=1 i j j ∂xj i,j,k=1 xi wij xj wjk xk
Pn = Pn
i=1 xi i=1 xi
Pn S S
Pn A A
i,j,k=1 xi wij xj wjk xk i,j,k=1 xi wij xj wjk xk
= Pn + Pn
i=1 xi i=1 xi
Pn P 2 P P 2
n S n n A
j=1 jx x w
i=1 i ij x
j=1 j x w
i=1 i ij
= Pn − Pn . (12)
i=1 xi i=1 xi
Hence, a symmtric interaction between subpopulations i and j increases the r, and an anti-
symmetric interaction between subpopulations i and j decreases the r. Competition and symbiosis
are the former type, and predator and prey are the latter type.
The law of mass action from chemical reaction theory states that a chemcial reaction like
k
n1 X1 + n2 X2 + · · · nν Xν −→ m1 Y1 + m2 Y2 + · · · mµ Xµ (14)
has a rate constant k, and the rate of reaction J, e.g., number of chemical reaction (14) per unit
time:
J = kxn1 1 xn2 2 · · · xnν ν ,
where xk is the concentration of chemical species Xk among the reactants. Then
dxk
= −nk J, k = 1, 2, · · · , ν;
dt
and
dyk
= mk J, k = 1, 2, · · · , µ,
dt
where yk is the concentration of chemical species Yk among the products.
We observe that dc dt
+ de
dt
= 0 and dsdt
+ dc
dt
+ dp
dt
= 0. This can be understood by going
through the biochemical reaction “mechanism” and recognize that the total enzyme e0 and
total substrates s0 are conserved. Substituting equations
c + e = e 0 , s + c + p = s0
Non-dimensionalization. The two equations in (18) are not yet ready to be analyzed computationally.
Note that since a computation has to have all the parameters in the equations assigned with
numerical values, explore the general behavior of a system differential equations involves many
calculations for differet parameter values. Thus the fewer the parameter, the better. The system
(18) seems to have four parameters: k1 , k−1 , k2 , e0 , and s0 . But actually, it has less.
Note that the one does not have to use the standard unit, such as molar, for the concentrations
s and c, nor standard unit for time, such as second, for t. Rather, one can try to use some
“internal units”. First, we note that k−1 must have “dimension” of [time]−1 since k−1 c ∼ ds dt
which is [concentration][time]−1 . Similarly, k1 has a dimension of [concentration]−1 [time]−1 ,
thus k1 e0 has a dimension of [time]−1 . Now let us introduce “non-dimensionalized variables”
s c
u= , v = , and τ = k1 e0 t (17)
s0 e0
in which combined parameters e0 /s0 = ϵ, k2 /(k1 s0 ) = λ and (k−1 + k2 )/(k1 s0 ) = K are all
dimensionless. We finally arrive at
du
= −u + (u + K − λ)v, (19a)
dτ
dv
ϵ = u − u + K v, (19b)
dτ
u(0) = 1, v(0) = 0. (19c)
in which we have denoted E[X] by µ. Two most important examples of random variables
taking real values are “exponential” and “normal”, also called Gaussian.
The pdf of a function of a random variable X. Let us have a random variable X with pdf
fX (x). Now consider a differentiable, monotonic increasing function g(·) and let Y = g(X).
So Y is also a random variable. What is the distribution of Y ? We note that
h i
Pr{Y < y} = Pr{X < g −1 (y)}, i.e., FY (y) = FX g −1 (y) . (25)
Therefore,
Z g −1 (y)
d d d
fX (x)dx = fX g −1 (y) g −1 (y) .
fY (y) = Pr Y < y = (26)
dy dy −∞ dy
which reads “fT (t)dt is the probability of random time T being in the interval (t, t + dt]. Then,
at time t, the probability the atom is still no decayed, i.e., T > t, is the survival probability:
Z ∞
p(t) = Pr{T > t} = fT (s)ds. (32)
t
We therefore have
dp(t)
fT (t) = − = re−rt . (33)
dt
The random time T has an exponential distribution. Its mean value, also called expected
value, is Z ∞
1
⟨T ⟩ = tfT (t)dt = . (34)
0 r
In fact, there is a variance in the random time T :
2
2 2 1
V ar[T ] = ⟨T ⟩ − ⟨T ⟩ = . (35)
r
T ∗ = min T1 , T2 .
(36)
And we have
Pr{T ∗ > t} = Pr{T1 > t, T2 > t} = Pr{T1 > t} Pr{T2 > t}. (37)
This is because the multiplication rule of two independent random events: The joint probability
is the product of the probabilities. Therefore, if one has n identical and independently distributed
random times T1 , T2 , · · · , Tn , then their minimum T ∗ has a distribution
n
Pr T ∗ > t = Pr{T1 > t} · · · Pr{Tn > t} = φT (t) ,
(38)
in which φT (t) = Pr{T > t} is a monotonically decreasing function with φT (0) = 1 and
φT (∞) = 0. Therefore, if φ′T (0) = r ̸= 0 and n is very large, we have
n n
t ′ t
lim φT = lim 1 + φT (0) = e−rt . (39)
n→∞ n n→∞ n
Why is there a 1/n on the left-hand-side of Eq. (39)? This is because with larger and larger
n, the mean time for T ∗ is getting smaller and smaller. In fact, it scales as 1/n. If we had not
n
introduced the 1/n, the limit of φT (t) would be 0 for all t > 0.
n exponential iid. In statistics, “iid” stands for “identical and independently distributed”.
If we consider n idential, independent atoms, each with an exponential waiting time e−rt , then
the time for the first decay, T ∗ = min{Ti , 1 ≤ i ≤ n} follows the distribution
n
Pr{T ∗ > t} = Pr{T1 > t, · · · , Tn > t} = Pr{T > t} = e−nrt . (40)
Note we have used the fact that all Ti are independent. Therefore, the rate for one decay from
n atoms is nr.
Exponential time is memoryless. Two measurements of T , one starts at t = 0, another
starts at t = t0 , will give identical result:
Pr{T > t0 + t} e−r(t0 +t)
= = e−rt . (41)
Pr{T > t0 } e−rt0
In an infinitesimal time interval (t, t + dt], the change in the survival probability of a single
atom is rp(t)dt.
Now consider a population of identical, independently distributed (iid) atoms. Let pn (t) be
the probability of having n radioactive atoms. There are two events that change the pn (t):
(a) A decay of one of n+1 radioactive atoms. This increases pn (t) while decreases
pn+1 (t); the rate is (n + 1)r.
(b) A decay of one of n radioactive atoms. This decreases pn (t) while increases
pn − 1(t). The rate is nr.
Therefore, considering each event can ocurr in the infinitesimal time interval (t, t + dt], we
have
dpn (t) = (n + 1)rpn+1 (t)dt − nrpn (t)dt. (43)
We now consider a population with N total individuals at t = 0. The individuals are
identical and independent, with individual “death rate”, i.e., death rate per capita, r.
To characterize the dynamics of population, X(t), X takes values 0, 1, 2, · · · , N , one no
longer can say that at time t, the X(t) is such and such. However, one can predict at time t, the
probabilitity of X(t) = n:
pn (t) = Pr{X(t) = n}. (44)
The pn (t) satisfies the system of differential equations
d
pn (t) = r(n + 1)pn+1 (t) − rnpn (t). (45)
dt
Then we have
d X dpn (t)
⟨X(t)⟩ = n
dt n=0
dt
X
= n (r(n + 1)pn+1 (t) − rnpn (t))
n=0
X X
= r n(n + 1)pn+1 (t) − r n2 pn (t)
n=0 n=0
X X X
= r (n + 1)2 pn+1 (t) − r (n + 1)pn+1 (t) − r n2 pn (t)
n=0 n=0 n=0
X
= −r (n + 1)pn+1 (t)
n=0
= −r⟨X(t)⟩.
8 Discrete-time dynamics
Not all dynamics requite a continuous counting of time. In fact, any realistic measurements
of any biological phenomenon are in discrete time. We now concern ourselves with dynamics
with discrete time. For population dynamics without immigration, these dynamics has the form
Nt+1 = Nt F Nt = f Nt . (47)
The simplest example of such dynamics is a linear system with F (N ) = a constant; the best-
known example of such nonlinear dynamics is the logistic growth with F (N ) = r 1 − N/K .
Interestingly, this is not really the discrete-time counterpart of the logistic differential equation.
r
A more faithful discrete time version of logistic differential equation is Fe(N ) = 1+N/K .
Now if we compare such dynamics with an ordinary differential equation (ODE) dx/dt =
f (x), and remember that one can study the ODE in terms of its distribution:
∂ρ(x, t) ∂
=− f (x)ρ(x, t) , (48)
∂t ∂x
then one expects that there is also “another equation” for the same dynamics in Eqn. (47). Note
that Eqn. (48) is a map of ρ(x, t) to ρ(x, t + dt), which is interpreted as distribution changing
with time. Therefore, we similarly have
Z ∞
ρ(y, t + 1) = ρ(x, t)K(x, y)dx. (49)
−∞
For each y, if K(y, x) is only concentrated at one point, then we say the dynamics is “deterministic”.
If there is a spread, we say the dynamics is “stochastic”. if for a given x to which therer are
more than one y, then we say the dynamics is “many-to-one”.
The most important properties of K(x, y) are ≥ 0, and
Z ∞
K(x, y)dy = 1 ∀x. (50)
−∞
A linear dynamics is “one-to-one”. The logistic map is “two-to-one”, but the Fe(N ) is one-
to-one. A stochastic dynamics can be “one-to-many”. One of the most important features of
the “1-to-1” dynamics is that one knows exactly where the dynamics is coming from and where
it goes. All other cases, there are some uncertainties, either in the past or in the future.
Note that this is called a “functional” with “al” at the end: It is a function of a function: For
each function ρ(x, t), Eqn. (52) returns a single scalar number. The ρ∗ (x) is considered known.
We now first show that H ρ(x, t) ≥ 0 for any normalized ρ(x, t), ρ∗ (x) ≥ 0:
Z ∞
ρ(x, t)
H ρ(x, t) = ρ(x, t) ln dx
−∞ ρ∗ (x)
Z ∞ ∗
ρ (x)
= − ρ(x, t) ln dx
−∞ ρ(x, t)
Z ∞ ∗
ρ (x)
≥ − ρ(x, t) − 1 dx
−∞ ρ(x, t)
Z ∞
≥ − ρ∗ (x) − ρ(x, t) dx
−∞
Z ∞ Z ∞
∗
≥ − ρ (x)dx + ρ(x, t)dx = 0.
−∞ −∞
More importantly even we don’t have ρ∗ (x), let us consider two sequences of ρ(x, t) and
ρ̂(x, t), started respectively with normalized ρ(x, 0) and ρ̂(x, 0)
h i Z ∞
ρ(x, t)
H ρ(x, t) ρ̂(x, t) = ρ(x, t) ln dx ≥ 0. (54)
−∞ ρ̂(x, t)
Now we consider
h i h i
H ρ(x, t + 1) ρ̂(x, t + 1) − H ρ(x, t) ρ̂(x, t)
Z ∞ Z ∞
ρ(y, t + 1) ρ(x, t)
= ρ(y, t + 1) ln dy − ρ(x, t) ln dx
−∞ ρ̂(y, t + 1) −∞ ρ̂(x, t)
Z ∞ Z ∞
ρ(y, t + 1) ρ(x, t)
= dx ρ(x, t) dyK(x, y) ln − ln
−∞ −∞ ρ̂(y, t + 1) ρ̂(x, t)
Z ∞ Z ∞
ρ(y, t + 1) ρ(x, t)
= dx ρ(x, t) dyK(x, y) ln − ln
−∞ −∞ ρ̂(y, t + 1) ρ̂(x, t)
Z ∞ Z ∞
ρ(y, t + 1)ρ̂(x, t)
= dx ρ(x, t) dyK(x, y) ln
−∞ −∞ ρ̂(y, t + 1)ρ(x, t)
Z ∞ Z ∞
ρ(y, t + 1)ρ̂(x, t)
≤ dx ρ(x, t) dyK(x, y) −1
−∞ −∞ ρ̂(y, t + 1)ρ(x, t)
Z ∞ Z ∞
ρ(y, t + 1)ρ̂(x, t)
= dy ρ(x, t)K(x, y)dx −1
−∞ −∞ ρ̂(y, t + 1)ρ(x, t)
Z ∞ Z ∞ Z ∞ Z ∞
ρ(y, t + 1)
= dy ρ̂(x, t)K(x, y)dx − dy ρ(x, t)K(x, y)dx
−∞ ρ̂(y, t + 1) −∞ −∞ −∞
Z ∞ Z ∞ Z ∞
ρ(y, t + 1)
= dy ρ̂(y, t + 1) − dy ρ(x, t)K(x, y)dx
−∞ ρ̂(y, t + 1) −∞ −∞
Z ∞ Z ∞
= dy ρ(y, t + 1) − dy ρ(y, t + 1) = 1 − 1 = 0.
−∞ −∞
So we have shown that
h i h i
H ρ(x, t) ρ̂(x, t) − H ρ(x, t + 1) ρ̂(x, t + 1) ≤ 0. (55)
Now, if the dynamics is one-to-one, then one can introduce a K −1 (x, y) such that
Z ∞
ρ(y, t) = ρ(x, t + 1)K −1 (x, y)dx. (56)
−∞
Then, all the above mathematics can be repeated, and one has
h i h i
H ρ(x, t) ρ̂(x, t) − H ρ(x, t + 1) ρ̂(x, t + 1) ≥ 0. (57)
Now combining Eqns. (55) and (57), we have
h i h i
H ρ(x, t) ρ̂(x, t) − H ρ(x, t + 1) ρ̂(x, t + 1) = 0, (58)
or h i
H ρ(x, t) ρ̂(x, t) = const. (59)
What is the significance of this mathematical result? Especially to biological dynamics?
(i) the event occurrence is homogeneous in time, with number of events per unit
time being r. r is the rate of the occuring events.
(ii) the occurrences of the events in disjointed intervale [t1 , t2 ] and [t2 , t3 ] are
independent;
(ii) in an infinitesimal time interval [t, t + dt], the probability of two events occur
is negligible, i.e., on the order of o(dt).
Therefore,
P (t + dt) − P (t) = −rP (t)dt + o(dt), (61)
taking the limit dt → 0, we have
d
P (t) = −rP (t). (62)
dt
We note that decay of a block of radioactive material is not homogeneous in time.
in which uℓ and wℓ are the birth and death rates with population ℓ. They are not rate per capita.
They are the rates for increasing one individual and decrease one individual, respectively.
Let us consider the simplest case of with birth and death rates, per capita, b and d. Then one
has un = nb and wn = nd. Let X(t) be the population in numbers, and pn (t) = Pr{X(t) = n}
be the probability of having n individuals in the population at time t. Then
d
pn (t) = (n − 1)bpn−1 − (nb + nd)pn + (n + 1)dpn+1 , (n ≥ 0). (64)
dt
∞
!
d d X
⟨X(t)⟩ = npn (t)
dt dt n=0
∞
X
= (n − 1)2 bpn−1 − n2 (b + d)pn + (n + 1)2 dpn+1
n=0
X∞ ∞
X
+ (n − 1)bpn−1 − (n + 1)dpn+1
n=1 n=0
∞
X ∞
X
= (n − 1)bpn−1 − (n + 1)dpn+1
n=1 n=0
= (b − d)⟨X(t)⟩. (65)
Indeed, the dynamics for the mean ⟨X(t)⟩ depends only one the difference of b − d. However,
one can also compute the variance of X(t):
in which ∞
X
2
⟨X (t)⟩ = n2 pn (t). (67)
n=0
Then,
∞
d d X 2
⟨X 2 (t)⟩ = n pn (t)
dt dt n=0
∞
X
b n2 (n − 1)pn−1 − n3 pn + d n2 (n + 1)pn+1 − n2 pn
=
n=0
∞
X
b n2 (n − 1)pn−1 − n(n + 1)2 pn + (2n + 1)npn
=
n=0
+ d n2 (n + 1)pn+1 − n(n − 1)2 pn − (2n − 1)npn
∞
X
= [b(2n + 1)n − d(2n − 1)n] pn
n=0
d d 2
⟨X (t)⟩ − ⟨X(t)⟩2
V ar[X(t)] =
dt dt
= 2(b − d)⟨X 2 (t)⟩ + (b + d)⟨X(t)⟩ − 2⟨X(t)⟩(b − d)⟨X(t)⟩
d
V ar[X(t)] = 2(b − d)V ar[X(t)] + (b + d)⟨X(t)⟩, (70)
dt
is a linear, constant coefficient, inhomogeneous, first-order ordinary differential equation. Its
solution can be obtained using the procedure in Sec. 9.4. Therefore, the mean and the variance
of the population X(t) are
⟨X(t)⟩ = Xo e(b−d)t , (71)
b + d (b−d)t (b−d)t
V ar[X(t)] = Xo e e −1 . (72)
b−d
The relative variance
V ar[X(t)] 1 b+d
1 − e−(b−d)t .
2
= (73)
⟨X(t)⟩ Xo b−d
We see that for the same net growth rate r = b − d, larger the b + d, larger the variance. In
a realistic population dynamics, the different rates of birth and death, b and d, matter; not just
their difference r = b − d.
Therefore, using the result in (76), we have the waiting time for the next event being T ∗ ,
which is exponentially distributed with rate uk + wk , and the probability of increasing one
individual with probability uku+w
k
k
, and decreasing one individual with probability ukw+w
k
k
. Thus,
the expected value for the change of the number of individual is
uk wk uk − wk
(+1) + (−1) = . (77)
uk + w k uk + wk uk + wk
And the rate of the mean population change :
uk − wk
uk + wk = uk − wk . (78)
uk + wk
For an infinitely large population, the randonness will be gone, and the rate of population
d
change is the mean value. Therefore, dt ⟨k⟩ = uk − wk .
A′ (t)e−rt = g(t);
A′ (t) = g(t)ert ;
Z t
A(t) = g(s)ers ds;
0
We have Z t
−rt −rt
x(t) = x(0)e +e ers ξ(s) ds = x(0)e−rt . (83)
0
More interestingly,
t/δ Z kδ 2
X
−rs
V ar x(t) = V ar[ξ] e ds
k=1 (k−1)δ
t/δ 2
e−r(k−1)δ − e−rkδ
X
= V ar ξ
k=1
r
1 − e−rδ
1 − e−2rt V ar[ξ]
=
r2 (1 + e−rδ )
( δ
−2rt
2r
rδ ≪ 1
≈ 1−e V ar[ξ] 1
(84)
r2
rδ ≫ 1
12 √
q
V ar x(t)
1 − e−rδ 1 − e−2rt
p
V ar[ξ]
= . (85)
x(t) r2 (1 + e−rδ ) e−rt x(0)
And for large time rt ≫ 1, we have a stationary stochastic dynamics x(t) fluctuating around
x = 0 with variance
h i 1 − e−rδ
stationary
V ar x (t) = V ar[ξ]. (86)
r2 (1 + e−rδ )
0.3
-0.05
0.4
0.1
0.3
-0.15
0.2
0.0
0.1
-0.1
-0.25
0.0
-0.2
0 2 4 6 8 0 2 4 6 8 0 1 2 3 4 5 6
x x x
Figure 2: The right-hand-side of the differential equation in (89), f (x, r, q), with q = 10, and
r = 0.5, 0.6, and 0.35. We observe that they corresponding to respectively, in addition to a
steady state at x = 0, “three steady states”, “one steady state”, and “one, another steady state”.
The ordinary differential equation (ODE) according to the law of mass action is
dx
= k1 ax2 − k2 x3 + k3 b − k4 x, (91a)
dt
da
= −k1 ax2 + k2 x3 , (91b)
dt
db
= −k3 b + k4 x. (91c)
dt
We note that combining the two reversible reactions in (90) yields an overall transformation
between A and B: A F GGGG
BG B.
A closed biochemical system. What is the steady state of the biochemical dynamics in
(90)? Letting the right-hand-side of Eq. (91) to be zero, we have
This yields
x∗ k3 a∗ k2 a∗ k2 k3
∗
= , ∗
= , =⇒ ∗
= . (93)
b k4 x k1 b k1 k4
This is in fact well-known in chemistry: Neglecting all the intermediates:
k1 k4
GGGG
AF B GGGG
G ··· F BG B,
k2 k3
The equilibrium relations in (93) determine the ratio of equilibrium concentrations, but not
their actual values. They have to be determined by the initial concentrations of the participating
chemical species.
In a closed biochemical reaction system, no matter how many different biocheimical species
involved in how many complex biochemical reactions, in the long time the system will reach a
chemical equilibrium.
An open biochemical system. Now consider a single living cell, as those in a cell culture
in a biomedical laboratory, as a complex biochemical reaction system. The “A” and “B” in
(90) can be glucose (C6 H12 O6 ) and CO2 +H2 O. The X can be all the important biochemicals
inside a single cell: vitamins, proteins, and DNA. Then the most important aspect in a cell
culture is to constantly change the “cell culture medium”, that is to keep the A and B out of
their equilibrium.
In fact, from the stand point of cell biochemistry, it is reasonable to simply assume that the
concentrations of A and B are at some constant level of a and b, fixed; not changing with time
at all. Such a device in a biomedical laboratory is called a “chemostat”.
Let us now consider some numbers: If we have k1 = 3, k2 = 0.6, both in the unit of
(mM)−2 sec−1 , and k3 = 0.25, k4 = 2.95 both in the unit of sec−1 , then ([A]/[B])eq = 0.6 ×
1
0.25/(3 × 2.95) = 59 . Fig. 3 shows the steady state of the biochemical system with fixed
concentrations of a and b for A and B.
3 * 1 * x^2 - 0.6 * x^3 + 0.25 * 1 - 2.95 * x
3 * 1 * x^2 - 0.6 * x^3 + 0.25 * 59 - 2.95 * x
0.0
2
10
0
-10
-1
-20
-2
x x x
Figure 3: The right-hand-side of the differential equation in (91a) with various fixed values of
a and b. Left panel: a = 1 and b = 59 give an equilibrium steady state in which xeq = 5.
Middle panel: a = 1 and b = 1 yield a bistable system with two stable steady states. Right
panel: a = 0.8 and b = 1, again a single steady state; the another one this time.
Then, a stable steady state of the ODE is represented by a minimum of the U (x), and an
unstable steady state of the ODE is represented by a maximum of the U (x). The dynamics
described by the differential equation can be visualized as a “down-hill” movement on a
“energy landscape”.
5
8
4
3
6
2
4
1
2
0
0
0 10 20 30 40 0 1 2 3 4 5
time time
Figure 4: Let panel: solutions to the ODE in Eq. (89) with q = 10 and r = 0.5; right panel:
solutions to the ODE in (91a) with a = b = 1, and other parameters k1 = 3, k2 = 0.6, k3 =
0.25, k4 = 2.95. Both ecological dynamics and biochemical dynamics exhibit bistability:
Depending on the initial state of a system, its ultimate fates can be very different. The unstable
steady state is often called a “threshold”.
Both Figs. (2) and (3) show that the number of steady states of a differential equation can
change with the parameters. One can in fact plot the steady state(s) of an ODE as a multi-valued
function of a parameter.
A multivalued scalar function usually is defined as the roots to an algebraic equation with
a parameter or several parameters, like f (u; q) = 0 or f (u; q, r) = 0. Assuming the f (u; q)
is smooth and differentiable with respect to both u and q, then in calculus we have learned
that the root to equation f (u; q) = 0 is a continuous curve in the (u, q) plane. Taking q as the
abscissa and u as the ordinate, this curve in general can have zero, one, or more u values for
each q.
Single variable f (u; q) = 0 ⇒ u(q). If f (u; q) is linear function of u, then:
4
2
0
parameter r
Figure 5: The steady states of the nonlinear ODE (89), which are the roots of its rhs f (x, r, q) =
x2
rx(1 − xq ) − 1+x 2 = 0, with q = 10 and r ∈ [0.35, 0.65]. It shows that the ODE can have either
1, or 2, or 3 steady states depending upon the value of r. Such a plot is called a bifurcation
diagram.
> library(rootSolve)
> uniroot.all(function(x) 0.5*x*(1-x/10)-xˆ2/(1+xˆ2),
lower=-1,upper=8)
yields
which should be compared with the left panel of Fig. 2. Now, by using a for-loop, we can have
Fig. 5 shows the multiple values of the roots to algebraic equation rx(1−x/10)−x2 /(1+x2 ) =
0 for the value of r ∈ [0.35, 0.65]. This figure should be compared with all three panels in Fig.
2. One of the most striking features of Fig. 5 is the “abrupt appearance or disappearance” of
a pair of roots, “out of blue”. This corresponds to a pair of roots “becoming complex” so they
no longer exist in the real space with x ∈ R.
Two-variable f (u; α, β) = 0 ⇒ u(α, β). We note that the rhs of Eq. (89) has actually
two parameters r and q. Therefore, the roots of f (u; q, r) are actually a scalar, multivalued
function of two independent variables. The curve in Fig. 5 then becomes a multi-layered
surface in a three-dimensional real space. Searching the words “catastrophe” with “Rene
Thom” on the web and looking for images, your will see how such a surface has a very novel
feature: Treating q and r as independent variables and u(q, r) as a multi-layered surface, there
are regions in (q, r) plane that correspond to a single layer of u, while other regions that have
three layers. At the boundary of these two regions the u has exactly two values.
One would like to be able to locate this boundary. Let us now solve this very intriguing math
problem. It requires some skill in your calculus. Using again the rhs of (89) as an example.
We already knew that it always has a root x = 0. So the remaining problem is to find the other,
possibly three, roots from
u
f (u; α, β) = α(β − u) − = 0, (96)
1 + u2
in which α = r/q and β = q. This change of notations simplifies a little bit of the algebra. Fig.
6A shows the root of the equation as a function of β, for several different α.
15 25
(A) (B)
12 20
roots u single root
9 15
6 10
3 5
single root
0 0
0 10 20 30 0 0.04 0.08 0.12
The situation with exactly two roots is a critical case. It occurs when f (u; α, β) is tangent
to the f = 0 axis, say at x = ξ. So both f (ξ) = 0 and f ′ (ξ) = 0 at ξ:
ξ 1 − ξ2
α(β − ξ) − = 0, −α − = 0. (97)
1 + ξ2 (1 + ξ 2 )2
If we eliminate the ξ from this pair of equations, we establish a relation between α and β,
which gives the boundary for the region in which the system has three roots.
Unfortunately, the elimination of ξ from Eq. (97) is not a simple task! However, we note
that we can obtain the following two equations from Eq. (97)
ξ2 − 1 2ξ 3
α= , β = , 1 ≤ ξ ≤ ∞. (98)
(1 + ξ 2 )2 ξ2 − 1
Recall that if both α and β can be expressed in terms of a parameter ξ, then Eq. (98) is
known as a parametric equation for the curve β(α). A well-known example is x = R cos t
and y = R sin t actually define a circle x2 + y 2 = R2 . Fig. 6B shows the function α vs. β: α
√ √ √
increases with ξ for ξ ∈ 1, 3 , then decreases with ξ > 3. There is a cusp at ξ = 3 .
One can understand the cusp qualitatively by simply considering the multi-layered surface
u(α, β) defined by Eq. (96).
pair of roots, “out of blue”. This corresponds to a pair of roots “becoming complex” so they
no longer exist in the real space with x ∈ R. the ODE in (91a) with constant a and b.
In nonlinear dynamical systems theory, the phenomenon of “abrupt appearance or disappearance”
of a pair of steady states, “out of blue”, is called a “saddle-node bifurcation”. It indicates
certain qualitative change in the dynamics. A plot of steady states as a multi-valued function
of a parameter, such as shown in Figs. 5 and 6A, are called bifurcation diagram. Then in the
case of two parameters, the behavior of the red, orange and green curves in Fig. 6A is known
as catastrophe. It involves two saddle-node bifurcation events.
Saddle-node, transcritical, and pitchfork bifurcations. The canonical forms are
√
dx 2 ss − µ non-existent when µ ≤ 0, unstable when µ ≥ 0
=µ−x , ⇒ x = √ (99)
dt µ non-existent when µ ≤ 0, stable when µ ≥ 0
dx 2 ss 0 stable when µ < 0 and unstable when µ > 0
= µx − x , ⇒ x = (100)
dt µ unstable when µ < 0 and stable when µ > 0
and
dx 0 stable when µ < 0 and unstable when µ > 0
= µx − x3 , ⇒ xss = √ (101)
dt ± µ non-existent when µ < 0, stable when µ > 0
where the positive ϵ represent a very small rate of immigration. There is a transcritical bifurcation
in the first model at K = 0. The two steady states of the second model are
p
ss K ± K 2 + 4Kϵ/r
X1,2 = ,
2
o o o
Figure 7: Bifurcation diagrams and corresponding vector fields before, during, and after
bifurcations. An arrow along a line indicates the directions of a vector field, while an open
circle and a filled circle represent a unstable and a stable fixed point, respectively. (A) Saddle-
node (out-of-blue) bifurcation has a pair of stable and unstable fixed points simultaneously
appear. (B) Transcritical bifurcation does not change the number of fixed points, rather there
is a switch of stability. (C) Pitchfork bifurcation turns a stable fixed point into a unstable one
surrounded by a pair of stable fixed points. All bifurcations shown here are “local”, which
means that a vector field has an infinitesimal local change at the critical bifurcation point when
µ = 0.
in which the positive and negative branches no longer interset for any K value. The transcritical
bifurcation phenomenon disappeared! From a biological standpoint, of course, the negative
K and negative X ss have no meaning. But as we shall show later in stochastic population
dynamics, there is a real significance of ϵ > 0, no matter how small.
Waddington’s epigenetic landscape.
The word energy derives from Greek ϵνϵ́ργϵiαζ (energeia), which possibly appears
for the first time in the work of Aristotle in the 4th century BC.
The concept of energy emerged out of the idea of vis viva (living force), which
Leibniz defined as the product of the mass of an object and its velocity squared;
he believed that total vis viva was conserved. To account for slowing due to
friction, Leibniz theorized that thermal energy consisted of the random motion of
the constituent parts of matter, a view shared by Isaac Newton, although it would
be more than a century until this was generally accepted. In 1807, Thomas Young
was possibly the first to use the term ”energy” instead of vis viva, in its modern
sense. Gustave-Gaspard Coriolis described “kinetic energy” in 1829 in its modern
sense, and in 1853, William Rankine coined the term “potential energy”. It was
argued for some years whether energy was a substance (the caloric) or merely a
physical quantity, such as momentum.
It has to wait for Einstein’s theory that unifies energy and mass: E = mc2 .
Energy conservation to include heat. In Eq. 105, the force from − dU dx
is called conservative
since kinetic energy and potential energy can forever convert back-and-forth. This is not the
case if there is an energy dissipation due to frictional force. A frictional force is proportional
to the velocity of a moving object:
d2 x dU (x) dx
m 2
=− −η , (106)
dt dt dt
dx
the last term no the right-hand-side is a frictional force. It is equal to zero if velocity dt
= 0.
Now parallel to the deviation of Eq. 105, we now have
" #
2 2
d m dx dx
+ U (x) = −η . (107)
dt 2 dt dt
The right-hand-side is the instantaneous rate of heat energy produced, which is equal to the
rate of energy decreasing in the mechanical system. The total mechanical energy (= kinetic +
potential) is no longer conserved in this system with friction. However, counting the rate of
heat dQ
dt
:
" # " #
2 2
d m dx dQ d m dx
+ U (x) = − ⇔ + U (x) + Q = 0. (108)
dt 2 dt dt dt 2 dt
The total energy conservation, including mechanical and thermal, is again regained.
The standard way to solve this linear, constant coefficient equation (109) is to assume the
general solution with the form ert . Then we obtain the characteristic polynomial for r:
mr2 + ηr + k = 0, (110)
We see that that if η ̸= 0 (η has to be positive from the physical requirement), then with
increasing t, x(t) in (112) tends to zero.
However, depending on whether η 2 ≥ 4mk or η < 4mk, the x(t) approaches to zero either
p
monotonically or oscillatorily with frequence 4mk − η 2 . The latter corresponds to Eq. 110
having a pair of complex roots.
Heavily overdamped system. When η 2 ≫ 4mk, the mechanical system is called heavily
overdamped. In this case, one can approximate the two roots in (111). We use the important
formula
s s2
(1 + s)1/2 ≈ 1 + − + ··· (113)
2 8
for small s. Then
p p
−η ± η 2 − 4mk −η ± η 1 − 4mk/η 2
r1,2 = =
2m 2m
−η ± η 1 − 2mk/η 2 − 2m2 k 2 /η 4
≈
2m
≈ − kη
−k/η 1 + mk/η 2
(
= η
−η/m 1 − mk/η 2 − m2 k 2 /η 4 ≈ − m
Both r1 and r2 are negative. Since η 2 ≫ 4mk, |r2 | ≫ |r1 |. Therefore, an overdamped system
has a very rapid acceleration phase in which “inertia balancing friction”, e.g., mẍ = −η ẋ, and
a relatively slow motion in which “friction balances elasticity”, i.e., η ẋ = −kx.
Significantly underdamped system. What happens if η 2 ≪ 4mk? In this case, we have
p √ p
−η ± η 2 − 4mk −η ± i 4mk 1 − η 2 /(4mk)
r1,2 = =
2m 2m
r
η k
≈ − ±i .
2m m
p
We have a decaying oscillation with frequency ω = k/m and a much slower decaying rate
η/(2m) ≪ ω. On the fast time scale, the inertia balances the elasticity: mẍ = −kx, just like a
Harmonic oscillation without damping.
Figure 9: Upper pannel: A schematic overview of protein-ligand complex separation with the
AFM. Lower pannel: One-dimensional model. The position of the ligand will be denoted by
x.
d2 x dx
m = −Fint (x) + k(d − x) − η , (114)
dt dt
in which x is the distance between the center-of-mass of the ligand to the center-of-mass of
the protein, which is assumed to be fixed.11 m is the mass of the ligand, η is its frictional
coefficient in water, k(x − x0 ) represents the force exerted by the AFM cantilever, with d being
10
Florin, E.L., Moy, V.T. and Gaub, H.E. (1994) Adhesion between individual ligand receptor pair. Science
264 415–417.
11
This immediately gives the insight that the internal structure of the protein can change under the pulling.
But if our meassurement for x is precisely the distance between the center-of-masses, then it does not matter.
However, in real world experiments, this is nearly impossible. So there will be consequences.
3.5
3
(A)
Ligand position z
2.5
1.5
0.5
05
0.5 1 5
1.5 2 5
2.5 3 5
3.5
Figure 10: Mechanical equilibrium position of the ligand, z, as a function of δ, the position
of the base of the cantilever, with several different αs, the stiffness of the cantilever. z = 1 is
the equilibrium position of the ligand in the absent of the AFM force. Red: α = 0.1; blue:
α = 0.3, and green: α = 0.7.
the position of the base of the cantilever. Fint (x) is the interaction force between the ligand
and the protein, it has the celebrated van der Waals potential Uvdw (x)
dUvdw (x) x0 6 x0 12
Fint (x) = , Uvdw (x) = −U0 2 − . (115)
dx x x
Because water is a rather viscous medium, we further assume that (1) the mechanical
system is overdamped, i.e., we can neglect the mass term. Therefore, Eq. (114) can be
simplified into
dx
η = −Fint (x) + k(d − x). (116)
dt
We now ask the question: When d is slowly increased, i.e., the AFM is pulling the ligand
away from the protein, how does the position of the ligand change?
This is in fact a static, force balance problem: Fint (x) = k(d − x). That is,
x 13
U0 x0 7 0
12 − 12 = k(d − x). (117)
x0 x x
(B)
3
Utot(z) 1
‐1
0 1 2 3 4 5
Ligand‐protein distance z
Figure 11: Total mechanical energy, Utot (z), as a function of the ligand-protein (center of
masses) distance z for several different values of δs. Red: δ = 1.3, green: δ = 2.2, and orange:
δ = 3.3. All with α = 0.1, correspond to the red curve in Fig. 1.
then,
z −7 − z −13 = α (δ − z) . (118)
Note that all three quantities, z, δ, and α are dimensionless. non-dimensionalization is a
very useful way to simplify mathematical models without involving any approximation. It uses
the internal scales as units for physical quantities in a model.
This equation can not be solved in a closed form for z(δ). However, one can obtain a
parametric equation for the function:
√ − 16
1 ± 1 − 4ξ ξ 1
z= , δ=z+ , ξ ∈ −∞, . (119)
2 αz 4
Fig. 1 shows several z as functions of δ with different α’s. We see with increasing α, i.e., the
spring becoming more stiff, the “sluggish” behavior disappears.
One can also understand the behavior in the figure in terms of the “potential energy function”:
dx dUtot
η =− , (120)
dt dx
where
1 2 x0 6 x0 12 1
Utot (x) = Uvdw (x) + k(x − d) = −U0 2 − + k(x − d)2 . (121)
2 x x 2
In non-dimensionalized form, it is
" 12 #
6
Utot (z) 1 1
=− 2 − + 6α(z − δ)2 . (122)
U0 z z
Fig. 2 shows the total potential energy function Utot (z) for three different values of δ.
Vector field
4
3
Prey population
2
1
0
0 1 2 3 4
Predator population
Figure 12: Predator-prey dynamics, as described by the differential equation (124), with
various initial values and α = 1: Red: u = 1, v = 0.1, orange: u = 2, v = 0.2, blue:
u = v = 2, brown: u = v = 1.7, and green u = v = 1.2.
αu + v − ln (uα v) = C, (128)
where C is a constant of integration. Now we consider a two-variable function H(u, v) =
αu + v − ln (uα v). It can be shown that H(u, v) has its minimum at u = v = 1, and the surface
has a curvature matrix
2
∂ H ∂ 2H
α/u2
!
∂u2 ∂u∂v 0
∂ 2H ∂ 2H = , (129)
0 1/v 2
∂v∂u ∂v 2
which is positive definite. That means the surface H(u, v) is “bowl like”. the the solution in
Fig. 13 are the contour curves of H(u, v) = C.
developed in 1950s. The latter is a system of four ordinary differential equations for (V, n, m, h)(t).
In contrast, the ML model is
dV
C = −gCa m∗ (V )(V − VCa ) − gK w(t)(V − VK ) − gL (V − VL ), (130a)
dt
dw w − w∗ (V )
= − , (130b)
dt τw (V )
in whivh
∗ V − v1
m (V ) = 0.5 1 + tanh , (130c)
v2
∗ V − v3
w (V ) = 0.5 1 + tanh , (130d)
v4
∗ −1 V − v3
τw (V ) = τ cosh . (130e)
2v4
In physiological applications, this model was developed for the dynamics with an interplay
between calcium ions and potassium ions in muscles. Note that one can re-write the Eq. (130b)
as
dw
= αw (V )(1 − w) − βw (V )w,
dt
with
w∗ (V )
αw (V ) = and βw (V ) = τw−1 (V ).
τ (V )
The implicit assumption of using m∗ (V ) in (130a) rather than a dynamic equation for m(t) is
that calcium dynamics is extremely fast on the time scale considered in Eq. (130).
We shall analyzing the ML equations with the following two sets of parameters:
Table I.
Parameter C gCa gK gL Vca VK VL (τ ∗ )−1 Iext v1 v2 v3 v4
Set 1 20 4.4 8 2 120 -84 -60 0.04 90 -1.2 18 2 30
Set 2 20 5.5 8 2 120 -84 -60 0.22 90 -1.2 18 2 30
Vector field
0.5
Gating Variable w
0.4
0.3
0.2
0.1
Membrane Potential V
Figure 13: Phase portrait of the Morris-Lecar excitable membrane dynamics, described by the
system (130) with the first set of parameters in Table I. Red line is the nullcline for dw dt
= 0
dV
and the blue line is the nullcline for dt = 0. Their intersection is a stable fixed point, a spiral
as illustrated by the orange trajectory. The black trajectory also indicates there is a stable limit
cycle. Between the stable fixed point and stable limit cycle, there is a unstable limit cycle as
shown by the green trajectory. The green trajectory is obtained by solving the system (130)
with t → −∞.
After non-dimensionalization:
r r r r
k2 k2 k1 k2 k−3 k2
x= cX , y = cY , t = k3 τ, a = cA , b = cB ,
k3 k3 k3 k3 k3 k3
we have
dx dy
= a − xy 2 = f (x, y), = b − y + xy 2 = g(x, y). (133)
dt dt
Planar system (133) has a single, positive steady state:
a
x∗ = , y ∗ = a + b,
(1 + b)2
at which, the Jacobian matrix
∂f ∂f
−(y ∗ )2 −2x∗ y ∗
! !
∂x ∂y
A= ∂g ∂g
=
∂x ∂y
(y ∗ )2 −1 + 2x∗ y ∗
(x∗ ,y ∗ )
When tr(A) = 0, fixed point changes from stable to unstable. This is called a Hopf bifurcation.
The Jacobian matrix actually provides a “frequency” for the spiral. The two eigenvalues are
1 p
λ1,2 = tr(A) ± tr2 (A) − 4 det(A)
2
whose imaginary part, at the critical condition of Hopf bifurcation is det(A) = (a + b)2 .
g0
amino acids + DNA −→ TF + DNA, (137b)
g1
amino acids + DNA · TFm −→ TF + DNA · TFm , (137c)
d
TF −→ amino acids . (137d)
12
Crick, F. H. C. (1958) On protein synthesis. Symp. Soc. Exp. Biol. 12, 139–163.
If g0 < g1 , we say the gene expression has a positive feedback; if g0 > g1 , we say the gene
expression has a negative feedback.
In the simplest form, the mathematical model for the biochemical system in (137) is a
planar system. We use X to denote the probability of the DNA with mTF bound, thus (1 − X)
for the probability of the DNA without TF, and Y as the concentation of the TF:
dX dY
= αY m (1 − X) − βX, = g0 a(1 − X) + g1 aX − dY, (138)
dτ dτ
in which a stands collectively for the concentration of amino acids, which is assumed to be a
constant.
Now, with non-dimensionalization:
Y g0 β α m
x = X, y = , t = τ d, g = , ω = , θ = g1 a ,
g1 a g1 d β
we have
dx h
m
i dy
= ω θy (1 − x) − x = f (x, y), = g + (1 − g)x − y = h(x, y). (139)
dt dt
Very large ω ≫ 1. If the FT unbinding to DNA is much more rapid than its own
degradation, i.e., ω = βd ≫ 1, then x(t) reaches its quasi-steady state quickly while y barely
changes:
θy m
x(y) = .
1 + θy m
Therefore, substituting this into the second equation in (139),
dy g + θy m
= − y. (140)
dt 1 + θy m
Very small ω ≪ 1. If the FT unbinding to DNA is much slower than its own degradation,
i.e., ω = βd ≪ 1, then this time y(t) reaches its quasi-steady state quickly while x barely
changes:
y(x) = g + (1 − g)x.
Substututing this into the first equation in (139), we have
dx n im o
= ω θ g + (1 − g)x (1 − x) − x . (141)
dt
We shall use mi and pi for the concentrations of mRNA and protein of FT-i:
dmi α1
= f pi−1 − mi , f (p) = α0 + , (142a)
dt 1 + pn
dpi
= −β pi − mi , (142b)
dt
in which i = 1, 2, 3 and p0 = p3 . That is, the (i − 1)th protein inhibits the synthesis of
ith mRNA. This model is known as repressilator, e.g, repression-driven oscillator. It is a
successful stroy of several independent engineering studies in 2000: A single pair of (m, p)
developed by Becskei and Serrano, two pairs of (m, p) giving rise to bistability investigated by
Gardner, Cantor, and Collins, and three pairs of (m, p), as in an oscillatory system (142) by
Elowitz and Leibler.
The steady state of the three systems with one, two, and three pairs of (m, p) can be
obtained as the roots to
f f f (x) − x = 0, (143)
| {z }
k
in which k = 1, 2, 3. Note that f (x) is a monotonically decreasing function of x. Hence
f (f (f (x))) is also a monotonically decreasing function of x. This implies there is only a
single root to Eq. (143). It is the same root as f (x) = x:
x − α0 1 + xn − α1 = 0.
(144)
On the other hand, the function f (f (x)) is actually a monotonic increasing function of x.
4
f(f(f(x))) f(f(x))
3
2 f(x)
0
0 1 2 3 4 5
x
Figure 14: f (x) = x, f (f (x)) = x, and f (f (f (x))) = x all have a same root. However,
f (f (x)) = x has two additional roots. Parameters α = 4 and m = 2.
Fig. 14 shows that f (x), f (f (x)) and f (f (f (x))) all intersect with x at a same x∗ . The
system with two pairs of (m, p) are two genes with mutual repression. It actually constitutes a
positve feedback, as shown by the red curve in Fig. 14.
characterization A = A ⃗h, α .
In many ecological problems, however, the complete list of conserved quantities is difficult
to obtain. In this case, a changing of perspective from extensive quantity to intensive quantity
solves the problem of non-ergodicity. Coming with this change, however, is an introduction of
uncertainty: The theory of probability enters into the deterministic dynamical systems theory.
in which
H1 (x, y) = α x − ln x + y − ln y . (145b)
There is a very important, distinct feature in this system: A given Hn can in fact correspond
to many different possibility of H1 , H2 , · · · , each one of them are conserved in the dynamics.
Such dynamical system is known as non-ergodic. Therefore, treating n as an external variable,
the meaning of
An+1 − An
h i
= Hn+1 x1 , y1 , x2 , y2 , · · · , xn , yn , xn+1 , yn+1 − Hn x1 , y1 , x2 , y2 , · · · , xn , yn
h,α
(146)
requires a careful analysis. One way to carry out this analysis is to change from a constant h,
an extensive quantity, perspective to a constant θ, an intensive quantity, perspective.
µX = µoX + kB T ln cX .
It has two parts: the first part, µoX , is solely determined by the nature of chemical structure of
a chemical species. It is related to something called “the internal energy”. The second part
is related to the amount of the chemical in a system. kB is kown as the Boltzmann constant:
1.3806488 × 10−23 m2 kg s−2 K−1 , T is temperature in Kelvin.
The chemical potential of the right-hand-side of reaction (147) is µA +µB ; and the chemical
potential of the right-hand-side of the reaction is µC + µD . In a chemcial equilibrium, one has
µA + µB = µC + µD . (149)
This leads to
ceq eq
C cD
µoA + µoB − µoC − µoD = kB T ln . (150)
ceq eq
A cB
From the dynamic equation in (148), however, we have
eq eq
cC cD k+
eq eq = . (151)
cA cB k−
Putting these together, we have the chemical potential difference across the reaction (147)
k+ cA cB
∆µ = µA + µB − µC − µD = kB T ln . (152)
k− cC cD
Now consider a chemical reaction as (147) in a controlled test tube, where all the four
chemcial species are activated being maintained by an experimenter. Then The reaction flux,
i.e., the net number of reactions per unit time, from the left to the right, is
J = k+ cA cB − k− cC cD . (153)
biology, there are three major “energy sinks” at the cellular level: (i) biosynthesis, (ii) powering
mechanical movements, and (iii) sustaining ionic and chemical gradients. Note that these three
ways of using energy are very classic; well established in 18th and 19th centuries. How about
“information processing”? Does information processing require energy expenditure?
In current cell biology, information processing is known as “regulations” and “signalings”.
Reversible enzyme kinetics. Let us again consider an enzmatic reaction:
k̂+1 k+2
GGGG
E+S F B GGGG
G ES F BG E + P. (157)
k−1 k̂−2
Now consider this is a single-enzyme system in terms of Markov probability p0 (t) and p1 (t)
for the states E, and ES at time t:
dp0
= k+2 p1 − k−2 + k+1 p0 + k−1 p1 , (158a)
dt
dp1
= k+1 p0 − k−1 + k+2 p1 + k−2 p0 , (158b)
dt
in which we introduced two new notations k+1 = kˆ1 cS and k−2 = k̂−2 cP . cA and cB are
assumed to be constant in a living steady state.
Now let us solve the steady state probabilities pss ss
0 and p1 from (158), and more importantly
the steady state flux from S → P :
k−1 + k+2
pss
0 = ,
k+1 + k−1 + k+2 + k−2
k+1 + k−2
pss
1 = ,
k+1 + k−1 + k+2 + k−2
ss
JS→P = pss ss ss ss
0 k+1 − p1 k−1 = p1 k+2 − p0 k−2
ss Vf cS
JS→P = .
K M S + cS
Now consider this is a single-enzyme system in terms of Markov probability p0 (t), p1 (t) and
p2 (t) for the states E, ES1 and ES2 at time t:
dp0
= k+3 p2 − k−3 + k+1 p0 + k−1 p1 , (162a)
dt
dp1
= k+1 p0 − k−1 + k+2 p1 + k−2 p2 , (162b)
dt
dp2
= k+2 p1 − k−2 + k+3 p2 + k−3 p0 , (162c)
dt
in which we introduced two new notations k+1 = kˆ1 cA and k−3 = k̂−3 cB . cA and cB are
assumed to be constant in a living steady state.
Now let us solve the steady state probabilities pss ss ss
0 , p1 , and p2 from (162), and more
importantly the steady state flux from A → B:
α2
GGGG
E∗ + P F BG E + P i + P, (165b)
β2
in which E ∗ is the phosphorylated form of enzyme E; K stands for a protein kinase, and P
stands for a phosphatase.
If one combines the two reactions in (165), then
α1 β2
GGGG
AT P F B GGGG
G ··· F BG ADP + P i,
β1 α2
in which parameters
α1 cAT P β 2 cP i
θ1 = and θ2 = .
α2 α2
Fig. 15 shows the fraction of phosphorylated E ∗ as a function of θ1 (cK /cP ) with various
values of ATP hydrolysis ∆µ. With small ∆µ, the upsteam kinase can no longer signal the
phosphorylation the down-stream substrate enzyme.
1
down‐steam phosphorylation
0.8
0.6
0.4
0.2
0
0.01 0.1 1 10 100
up‐steam kinase activity
We now introduce continuous variable x = n/b, and similarly m/b = z. Then the sum in Eq.
(169) can be written as an integral, through a Riemann integral: patition, sum, taking limit.
First, let us denote
ubz−1 u(z)
lim = . (170)
b→∞ wbz w(z)
Then the sum n Z x
X um−1 u(z) 1
ln =b dx ln , dx = . (171)
m=1
w m 0 w(z) b
Now let us consider birth and death rates according to the ecological model given in Eq.
(87):
rn2 an2
un = rn, wn = + 2 . (172)
q b + n2
Then
u(x) rbx
= lim r(bx)2 a(bx)2
w(x) b→∞ +
q b2 +(bx)2
rb αβ
= lim 2 = lim x (173)
b→∞ rb x ax
+ 1+x2 b→∞ αx +
q 1+x2
0 3 0 3 6 9 12 0 3 6 9 12
0.8 0.4 0.2
=10 0.2 =12.7
04
0.4 0
0 =12
0 ‐0.2
‐0.2
‐0.4 ‐0.4 ‐0.4
0 10 20 0 10 20 30
(x) 0
=14
14
0
=19
19
‐0.5 ‐2
‐1 ‐4
x ‐1.5
15 ‐6
6
Figure 16: ϕ(x) given by Eq. (175) with different values of β and α = 0.03, corresponding to
σ = 5.86.
1+α
in which σ 2 = α
A is a constant, and
,
2
σ + x2
x
x
ϕ(x) = x ln + 2σ arctan + x ln − 2 arctan x − x. (175)
1 + x2 σ β
15.2 Relation between deterministic and stochastic steady states and time
scales
Fig. 16 shows that ϕ(x) for α = 0.03 and β = 10, 12, 12.7, 14 and 19. Comparing Figs. 2
and 16, it seems that the minima of ϕ(x) are located at the steady state of ordinary differential
equation (89). This turns out to be exactly true: The minima of ϕ(x) are located exactly at the
stable steady states of the ODE; and the maxima are located at the unstable steady states of the
ODE. To show that we only have to carry out the derivative
dϕ(x) αβ(1 + x2 ) b(x)
= − ln 2
= − ln . (176)
dx αx(1 + x ) + x d(x)
d
Therefore, steady states xs where b(xs ) = d(xs ) is also the place dx ϕ(xs ) = 0.
We note that with b → ∞, the probability density function f (x) → δ(x − x∗ ) where x∗ is
the global minimum of ϕ(x). The global minimum will have probability 1 while all the local
minima have probability 0. A local minimum is called a metastable state.
The concept of Lyapunov property. If for a deterministic dynamics ẋ = f (x) a function
L(x) satisfies
d
L (x(t)) ≤ 0, (177)
dt
Prof. Hong Qian 66 Thursday 5th January, 2023, 09:25
AMATH 423/523 Mathematical Analysis in Biology and Medicine Winter, 2023
then we say function L(x) has Lyapunov property with respect to the dynamics. ϕ(x) has
Lyapunov property with respect to the ODE ẋ = b(x) − d(x):
d dϕ(x) dx b(x)
ϕ (x(t)) = = − ln × (b(x) − d(x)) ≤ 0. (178)
dt dx dt d(x)
In an evolutionary time scale, ODE’s t = ∞ is very short. For any finite b, i.e., finite
population, its dynamics is stochastic and in a correspondingly long time, ∼ ecb (c > 0) the
dynamics will have finite probabilities near both stable steady state. This is represented by the
e−bϕ(x) . However, in order for such a stationary distribution to emerge, the dynamics has to be
in the two regines back and forth many times. This is a time scale beyond the infinite of the
ODE dynamics. We shall call this “evolutionary time scale”.
Some philosophical implications. Deterministic dynamics is the cornerstone of Newtonian
theory of our physical world. Simple differential equations have stable steady states. The
dynamics is usually “converging” to stable steady states; depending on the initial condition.
However, as we have shown, this coverging dynamic view is only valid on a very short
time scale. In a much longer “evolutionary” time scale, the dynamics will be “diverging”.
For highly nonlinear systems, there are many many stable attractors, and the “evolutionary
dynamics” is stochastic and jumps among all the different attractors. Deterministic dynamics
is intra-attractoral while the stochastic evolutionary dynamics is inter-attractoral. They are on
a completely different time scale.
dx
= rx(1 − x). (180)
dt
Linear growth rate and quadratic death rate. un = rn and wn = rn2 /q. Then we have
r(n + 1)2
d n
pn (t) = r(n − 1)pn−1 − rn 1 + pn + pn+1 . (181)
dt q q
This function has a single minimum at x = 1, corresponds to the stable steady state of Eq.
(180).
Pure birth process with decreasing birth rate. un = rn(1 − n/q) and wn = 0. Then,
d n−1 n
pn (t) = r(n − 1) 1 + pn−1 − rn 1 + pn . (183)
dt q q
The long time dynamics n = q is an absorbing state. The stationary distribution is psn = δn,q .
That is, x = 1 is an absorbing state of the system with stationary distribution f s (x) = δ(x − 1).
One can, however, compute the so-called quasi-stationary distribution, i.e., the distribution
among the population that has not been absorbed:
Note all three quantities, k+1 [S], k−1 , and k2 have the physical dimension of [time]−1 , as the
rates of exponentially distributed waiting time for stochastic elementary reactions.
Let ⟨T ⟩ be the mean time for the completion of a kinetic cycle, e.g., from the E on the
left to the E on the right of (185). We add the mean time for each and every step. First, the
−1
transition from E → SE has a mean time of k+1 [S] , this is followed by a waiting time in
the state SE, (k−1 + k2 )−1 . After that, there are two possibilities: the SE → S + E with a
k−1
probability k−1 +k2
and the SE → P + E with a probability k−1k+k 2
2
. Therefore, the mean time
k−1 ⟨T ⟩+k2 ·0
to produce a P , after leaving state SE, is k−1 +k2
. Puttng all these together, we have
1 1 k−1 ⟨T ⟩
⟨T ⟩ = + + , (186)
k+1 [S] k−1 + k2 k−1 + k2
from this one can solve ⟨T ⟩. Noting the the rate of the enzyme reaction is the reciprocal of the
mean time per product formation, we have
1 k−1 + k2 1
= ⟨T ⟩ = + , (187)
v k+1 k2 [S] k2
or,
k+1 k2 [S] Vmax [S]
v= = , (188)
k−1 + k2 + k+1 [S] KM + [S]
in which Michaelis constant KM = k−1 +k2
k+1
, and Vmax = k2 . In the standard theory of MM
enzyme kintics, the Vmax = k2 × enzyme concentration. This is precisely expected from many
enzymes working together with each individual molecule having the rate given in (188).
This stochastic formulation is the ultimate scenario of [E] ≪ [S].
in which b is the probability of infection after each bit, from a mosquito to a host, and the b̂ is
the probability of infection after each bit from an infectious host to a uninfected mosquito! a
is the mean number of bits per unit time, M is the population number of the mosquito, and γ
is recovery rate of the infectious person.
xk (t) is the population of the infected individuals in the subbroup k at time t. in which k ≡
PK
j=1 jd j . Note that for K = 1, this is the logistic growth model with growth rate λ − r and
r
carrying capacity d1 (1 − λ ).
Eq. 193 is a system of K coupled nonlinear ODEs. It is clear that the origin is a steady
state. To find the other steady state(s), let us intruduce the function
K
1X
C {xi } = ixi ≥ 0. (194)
k i=1
that is when
r K
P
j=1 jdj
λ > λ c ≡ PK . (198)
2
j=1 j dj
∗
rλkℓ dk
= − r + λkC δkℓ + . (200)
k r + λkC ∗
The fixed point at origin corresponds to C ∗ = 0. We thus have the linear matrix
λkℓdk
Bkℓ = −rδkℓ + . (201)
k
The matrix λkℓdk /k can be written as
0 ··· d1 d1 · · · 1 0 ···
1 0 d1 0
λ 0
2 ··· 0 d2 d2 · · ·
d2 0 2 · · ·
0
.. .. . . .. .. .. . . .. .. .. . . .. ,
k . . . . . . . . . . . .
0 0 ··· K dK dK · · · dK 0 0 ··· K
which has only a single non-zero eigenvalue. Therefore, the Bkℓ in (201) has eigenvalues −r
with multiplicity of (K − 1) and λk (d1 + 4d2 + · · · + K 2 dK ) − r. We see that when λ is greater
than the critical λc in (198), the origin becomes unstable. This coincides with the existence
of the non-zero steady state (199). The model predicts that when there is a breakout of an
epidemic, (0, 0, · · · , 0) is unstable and (x∗1 , x∗2 , · · · , x∗K ) exists: Each and every subgroup x∗j is
non-zero.
Let us now consider the linear stability of the steady state given in (199), with the corresponding
Jacobian matrix
1 0 ··· 0
0 2 ··· 0
Bkℓ = −rI − λC ∗ ... ... . . . ...
(202)
0 0 ... K
α1 α1 · · · α1 1 0 ··· 0
rλ α..2 α..2 ·.·. · α..2 0.. 2.. ·.·. · 0.. ,
+ . . .
k . . . . .
αK αK · · · αK 0 0 ··· K
in which
kdk
αk = .
r + λkC ∗
The matrix has eigenvalues −λC ∗ , −r − 2λC ∗ and −r − 3λC ∗ .
1 0 0 α1 α1 α1 1 0 0 0
0 2 0 α2 α2 α2 0 2 0 1
0 0 3 α3 α3 α3 0 0 3 0
1 0 0 α1
=2 0 2
0 α2
0 0 3 α3
17.2
dx1 λ
= −rx1 + d1 − x1 x1 + 2x2 ,
dt
k (203)
dx2 2λ
= −rx2 + d2 − x2 x1 + 2x2 .
dt k
is λ2 − (a + d)λ + (ad − bc) = 0, which has the discriminant being (a + d)2 − 4(ad − bc) =
(a − d)2 + 4bc. So a necessary, but not sufficient, condition for having complex eigenvalues is
bc < 0. Looking at the matrix in (205), we see that both eigenvalues are real: one of them is
−r, and the other is
λ
d1 + 4d2 − 2x1 − 8x2 − r.
k
17.5 Per capita infection rate: density vs. frequency dependence and
Michaelis-Menten kinetics
x
x xT
= K x
. (207)
K +x xT
+ xT
18 Numerical methods
18.1 Euler’s method
18.2 Runge-Kutta method
18.3 von Neumann rejection method for random number generation
13
18.4 Tau-leaping
18.5 First-reaction and next-reaction methods
14
in which
ξi = x − ci t + ϕi , (210b)
√
ci = 2 1 + β − 3λi , i = 1, 2, (210c)
β 1
λ1 = √ , λ2 = √ . (210d)
2 2
The solution is obtained as follows. Let us introduce a transformation
wx
u(x, t) = µ , (µ ̸= 0) (211)
w+σ
13
Chen, Y. (2005) Another look at rejection samplingthrough importance sampling. Statistics & Probability
Letters, 72, 277–283.
14
Gibson, M. A. and Bruck, J. (2000) Efficient exact stochastic simulation of chemical systems with many
species and many channels. Journal of Physical Chemistry A, 104, 1876–1889.
where σ is a constant. Substituting this into the original equation, we have, since constant σ is
arbitrary:
Let us consider the end point is N . This is famously known as “the gambler’s ruin problem”.
Then we have
TN = 0, and T0 = T1 . (218)
How do we solve the general solution for Tn ? Again, it is an inhomogeneous llinear
difference equation. The solution to the homogeneous problem is λn :
λn = qλn−1 + pλn+1 .
15
Petrovskii, S. and Li, B.-L. (2003) An exactly solvable model of population dynamics with density-dependent
migrations and the Allee effect. Mathematical Biosciences, 186, 79–91.
Note that if p = q = 12 , then the particular solution is −n2 . Therefore, the general solution
to (217) is n
q n
Tn = a1 + a2 + . (220)
p q−p
Applying the boundary conditions in (218) we have
N
N p q p
a1 = + , a2 = − .
p − q (p − q)2 p (p − q)2
Therefore, "
N n #
N −n p q q
Tn = + − , 0≤n≤N (221)
p−q (p − q)2 p p
Let us now discuss the solution in (221). First, if p > q, then for large n and N , the terms
−n
in the [· · · ] ≈ 0, and we have Tn ≈ Np−q — distance divided by the velocity. However, when
p < q:
N
p q
Tn→N ≈ ∼ eN ln(q/p)
(p − q)2 p
is actually independent of initial position n, and it is exponentially large with respect to N .
Catastrophe in bistable system is induced by a changing “environment”; but the rare events
in bistable system are spontaneous.