Open access peer-reviewed chapter

Dynamic Equations on Time Scales

Written By

Sabrina Streipert

Reviewed: 25 March 2022 Published: 15 March 2023

DOI: 10.5772/intechopen.104691

From the Edited Volume

Nonlinear Systems - Recent Developments and Advances

Edited by Bo Yang and Dušan Stipanović

Chapter metrics overview

158 Chapter Downloads

View Full Metrics

Abstract

An extension of differential equations to different underlying time domains are so called dynamic equations on time scales. Time scales calculus unifies the continuous and discrete calculus and extends it to any nonempty closed subset of the real numbers. Choosing the time scale to be the real numbers, a dynamic equation on time scales collapses to a differential equation, while the integer time scale transforms a dynamic equation into a difference equation. Dynamic equations on time scales allow the modeling of processes that are neither fully discrete nor fully continuous. This chapter provides a brief introduction to time scales and its applications by incorporating a selective collection of existing results.

Keywords

  • time scales
  • existence
  • uniqueness
  • linear
  • applications

1. Introduction

The modeling of processes using differential equations is a well-established method in multiple branches of sciences. Dependent on the model assumptions, the form of the differential equation can range from a comparably simple ordinary differential equation to more advanced formulations using nonlinear, higher order, and partial differential equations. Reasons to consider difference equations include computational benefits and, even more fundamental, a discrete modeling perspective. For example, when describing a zero-coupon bond where the invested amount at time t, Mt, receives interest r at the end of each year but remains unchanged during each year, the recursive model Mt+1=1+rMt captures the change of the investment from time t to time t+1. Difference equations are also a common tool to describe processes on a macro scale in time, for example, when describing non-overlapping generations. Even though the number of individuals may vary throughout the generation period, one may only be interested in the individuals at the beginning of each generation time, i.e., the size of each cohort. There are however processes that cannot be described accurately using differential or difference equations. For example, when modeling species that are reproducing continuously during certain months of the year before laying eggs right before hibernating. Another example are plant populations that grow continuously during some months of the year and plant their seeds prior to dying out. In [1], Robert May gives examples of insects that exhibit such hybrid continuous–discrete behavior.

Instead of introducing a set of simplifying assumptions and possibly discontinuous model parameters that impact the model analysis, dynamic equations on time scales can provide a simple alternative to describe such processes. Time scales calculus was introduced by Stefan Hilger in 1988 [2]. It unifies the continuous and discrete calculus and extends it to any nonempty closed subset of the real numbers called a time scale, denoted by T. By introducing differentiation and integration on T, the classical theory of differential equations can be extended to time scales, which allows the modeling of processes that are not changing continuously nor solely discretely in time. These so-called dynamic equations are essentially the time scales analogue of differential and difference equations and have gained increasing interest due to their potential in applications. Choosing the time scale to be the real numbers, a dynamic equation transforms into a differential equation and by choosing the time scale to be the integers, a corresponding difference equation is obtained. Thus, instead of studying differential equations and difference equations separately, time scales provides also a tool to investigate both by analyzing the corresponding dynamic equation. This is specifically interesting since certain difference equations exhibit significantly different behavior as their continuous analogues, see for example the “logistic map” and the “logistic differential equation”. By analyzing a dynamic equation on time scales, the effect of the underlying time domain onto the behavior of solutions may be revealed.

Advertisement

2. Time scales fundamentals

In this subsection, the basic definitions of time scales calculus are introduced based on the introductory book [3].

Definition 1. A time scale, denoted by T, is a nonempty closed subset of R.

Examples of a time scale are R,Z,hZ,qN0=1,q,q2,q3,q>1,abcd where a<b and a,b,c,dR, and the Cantor set. It therefore contains the popular cases of the continuous, the discrete, and the quantum calculus.

Operators that aid the description of a time scale are the “forward jump operator”, denoted by σt, the “backward jump operator”, denoted by ρt, and the “graininess function”, denoted by μt. These operators are defined for tT as

σtinfsT:s>t,ρtsupsT:s<t,μtσtt.E1

Since T is closed, σ,ρ:TT and μ:T[0,). Table 1 provides values of the corresponding operators for different examples of time scales.

Using these operators, any tT can be classified as:

  • right-scattered (left-scattered), if σt>t (ρt<t), and

  • right-dense (left-dense), if σt=t (ρt=t).

We say that a point tT is isolated, if it is right- and left-scattered. We say that a point tT is dense, if it is right- and left-dense. Note that for T=R, every point is dense and, for T=Z, every point is isolated.

Example 2.1. El Nin¯o events can be described using a time scale. El Nin¯o events between 2002 and 2017 have been observed in the time intervals 2002–2003, 2004–2005, 2006–2007, 2009–2010, and 2014–2016 [4], which suggests the corresponding time scale (Figure 1, Table 2)

Figure 1.

Part of the time line containing points in the time scale T. Curly lines identify intervals within T. Here, t120042005, t2 is the last point in the interval 20042005, and t3=2006 is the first point in 20062007.

T=i=05aiai+1

with a0a1a2a5=2002,2004,2006,2009,2014,2015.

The following notation is commonly used for tT,

σnt=σσσntimest,ρnt=ρρρntimest.

2.1 Functions on time scales

We can now consider scalar functions on time scales, that is, f:TR, and discuss their properties. We define the subset Tκ as follows: If T has a left-scattered maximum mT, then Tκ=T\m, else Tκ=T.

Definition 2. f:TR is called regressive, if, for all tTκ,

1+μtft0

and is called positively regressive, if, for all tTκ,

1+μtft>0.

The following are properties of f:TR that later identify integrability.

Definition 3. f:TR is called regulated provided its right-sided limit exists (as a finite value) for all right-dense points and its left-sided limit exists (as a finite value) for all left-dense points.

Even though every regulated function on a compact interval is bounded, in general, maxatbft and minatbft do not need to exist for regulated f:TR.

Definition 4. f:TR is called rd-continuous if f is continuous at all right-dense points and its left-sided limit exists (as a finite value) for all left-dense points. The set of rd-continuous functions is denoted by Crd=CrdT=CrdTR.

Note that, if f:TR is continuous, then f is rd-continuous. If f is rd-continuous, then f is regulated.

The set of rd-continuous and regressive (positively regressive) functions is denoted by R=RT=RTR (R+=R+T=R+TR).

Beside the classical addition and subtraction of functions, time scales calculus introduces the so-called “circle plus”, denoted by , and “circle minus”, denoted by . These operations are defined for f,g:TR as follows

fgt=ft+gt+μfgtand,forgR,fgt=ftgt1+μgt.

A useful property is that if f,gR (R+), then fg,fgR (R+) implying that the (positively) regressive property is being carried over. Furthermore, R forms an Abelian group with the inverse elements of fR given by f.

For T=R, the operators and correspond to the classical addition and subtraction.

2.2 Differentiation

Definition 5. Let f:TR and tTκ. If there exists fΔtR such that for all ε>0, there exists δ>0 such that

fσtfsfΔtσtsεσtsforallstδt+δT,

then we call fΔt the delta (or Hilger) derivative of f at tTκ.

If fΔt exists for all tTκ, we say that f is delta differentiable (or short: differentiable) and the function fΔ:TR is called delta derivative of f on Tκ.

If f is differentiable at tTκ, then

fσt=ft+μtfΔt.

The following notations are used equivalently

fσt=fσt=fσt.

The definition of a delta derivative can be extended to consider higher order derivatives. We say that f is twice delta differentiable with the second (delta) derivative fΔΔ, if fΔ is (delta) differentiable on Tκ2=Tκκ.

Note that the definition of delta derivatives focuses on the change forward in time. A corresponding definition that focuses on the change backward in time is referred to as nabla derivative, see for example [5].

Theorem 2.2. [See [3, Theorem 1.16]] Let f:TR and tTκ. Then, the following holds:

  1. If t is right-dense, then

    fΔt=limstftfsts,

    provided that the limit exists (as a finite number).

  2. If f is continuous at the right-scattered point t, then

fΔt=fσtftμt.

Applying Theorem 2.2 for the case of T=R, shows that the delta derivative is consistent with the classical derivative, that is, fΔt=ft for tT=R. For T=Z, the delta derivative collapses to the forward Euler operator, widely accepted as the discrete analogue of a derivative, that is, fΔt=ft+1ft if T=Z (see Table 3).

Tσtρtμt
Rtt0
Zt+1t11
qN0qttqtq1

Table 1.

The description of the time scales functions σ,ρ,μ for the examples of R, Z, and qN0(q>1).

tTσtμtρt
t120042005σt1=t1μt1=0ρt1=t1
t2=2005σt2=t3μt2=1ρt2=t2
t3=2006σt3=t3μt3=0ρt3=t2

Table 2.

The functions σ,ρ,μ for the time points t1,t2,t3T based on Figure 1.

TT=RT=ZT=q0
fΔtftΔftfqtfttq1

Table 3.

Derivatives for the examples of T=R, T=Z, and T=qN0 (q>1). Note that Δft=ft+1ft is the forward Euler operator.

As in the continuous case, the differential operator is linear, that is, for α,βR, tTκ, and for (delta) differentiable functions f,g:TR,

αf+βgΔt=αfΔt+βgΔt.

The analogues of the product and the quotient rule on time scales take on slightly different forms. For (delta) differentiable functions f,g:TR, and tTκ,

fgΔt=fΔtgσt+ftgΔt=fΔtgt+fσtgΔt

and, for gt,gσt0,

fgΔt=fΔtgtftgΔtgtgσt.

For T=R, we have fσ=f and gσ=g so that the classical product and quotient rule are retrieved. In the case of T=Z, we have the correspondent rules consistent with [6], namely

Δfgt=Δftgt+1+ftΔgt=Δftgt+ft+1Δgt.

If gt,gt+10, then

Δftgt=fgΔt=ΔftgtΔgtftgtgt+1.

The modifications in the product and quotient rule highlight that some of the well established differentiation rules only carry over to time scales calculus after some adjustments. In fact, the product rule on time scales reveals that the useful property of power functions ft=tn for nN0 is no longer the simple reduction of the power by one, because

t2Δ=ttΔ=t+σt,

which may not be delta differentiable. This indicates already that the series representation of functions requires further thought.

Also, considering the chain rule, we note that for T=Z,

Δfft=fσtfΔt+ftfΔt=fΔtft+fσt2ftfΔt,

for fσtft. Thus, the powerful chain rule, often utilized in solving differential equations via a variable transformation, does not apply on time scales. In an attempt to generalize the chain rule for functions on time scales a few identities have been formulated. The next theorem provides such an expression based on works in [7, 8]. Other formulations can be found in [3].

Theorem 2.3. (See [3, Theorem 1.90]). Let f:RR be continuously differentiable and suppose g:TR is (delta) differentiable. Then fg:TR is (delta) differentiable and

fgΔt=01fgt+tgΔtdhgΔt.

An interesting observation is that the operators, Δ and σ, do generally not commute, that is, fΔσfσΔ. Take for example T=qN0 with q>1, then

fΔσt=fq2tfqtμqtfq2tfqtμt=fσΔt,

since μqt=qtq1tq1=μt.

2.3 Integration

Definition 6. A continuous function f:TR is called pre-differentiable with (region of differentiation) D, provided that DTκ, Tκ\D is countable and contains no right-scattered elements of T, and f is (delta) differentiable at each tD.

Theorem 2.4. (See [3, Theorem 1.70]). Let f:TR be regulated. Then there exists a function F:TR which is pre-differentiable with region of differentiation D such that

FΔt=ftforalltD.

The function F is called an pre-antiderivative of ft.

If FΔt=ft for all tTκ, then F is called antiderivative of f.

We define the indefinite integral of a regulated function f by ftΔt=Ft+C, where CR is an arbitrary integration constant and F is a pre-antiderivative of f. The Cauchy integral is defined by abftΔt=FbFa for all a,bT.

Theorem 2.5. (See [3, Theorem 1.74]). Every rd-continuous function f has an antiderivative. In particular, if t0T, then F defined by

Ftt0tfsΔsforalltT

is an antiderivative of f.

For T=R, the integral is consistent with the Rieman integral (see Table 4).

TT=RT=ZT=qN0TI
stfτΔτstfτdττ=st1fτn=0ksqkq1fqksτabTμτfτ

Table 4.

Integrals for the examples of T=R, T=Z, and T=qN0 (q>1), and isolated time scales TI, for which all points in TI are assumed to be isolated. In all cases, s,tT and s<t. In the case of T=qN0, we assume t=qks.

The integral operator is linear so that for f,gCrd and a<b, a,bT, and α,βR,

abαf+βgsΔs=αabfsΔs+βabgsΔs.

With the definition of integration on time scales, we have the machinery to introduce a series representation for time scales functions. In [9], see also [3], a time scales analogue of polynomials that allows a corresponding Taylor series expression was introduced using the recursive formulation

g0ts=h0ts1forallt,sT,

and, for every kN0,

gk+1ts=stgkστsΔτforalls,tT,

and

hk+1ts=sthkτsΔτforalls,tT.

Now, hkΔts=hkts and gkΔts=gkσt,s for kN and t,s∈Tκ. Two Taylor series representations can be formulated for a time scales function f, one that uses the time scales polynomials gk and one that uses the polynomials hk, see Section 1.6 in [3] for more details.

Advertisement

3. Linear dynamic equations

This chapter provides a brief introduction to first order dynamic equations and provides a selected summary of [3], extended by applications. A first order dynamic equation is of the form

yΔt=ftyyσ,E2

for y:TRn and f:T×Rn×RnRn with nN1=1,2,3. A first order initial value problem (short: IVP) is then given by (2) with an initial condition yt0=y0Rn for t0T. A function y:TRn is called a solution of (2) if y satisfies the equation for all tTκ.

We call (2) linear if

ftyyσ=f1ty+f2t,orftyyσ=f1tyσ+f2t,

where f1,f2:TRn. We say the linear dynamic equation is homogeneous, if f20.

3.1 Scalar case

We first focus on the scalar case of (2), that is, f:TR. Based on the above definition of linearity, there are two forms a linear, homogeneous, first order dynamic equation can have:

yΔ=pty,E3
yΔ=ptyσ,forp:TRE4

Note that for T=R, yσ=y and therefore y=ptyσ=pty so that both, (3) and (4), are the time scales analogues of y=pty.

If pR, then (3) is called regressive and if pR, then (4) is called regressive.

The unique solution to (3) with initial condition yt0=1 for some t0T is denoted by yt=eptt0 and is called the time scales exponential function. The unique solution to (4) with initial condition yt0=1 is yt=eptt0.

Table 5 contains the time scales analogues of the exponential function for the dense time scale T=R, the discrete time scale T=Z, and the quantum time scale T=qN0.

The Table 5 reveals a crucial aspect of the time scales exponential function, namely that the positivity property, known for the traditional exponential function, does not uphold on time scales. Take for example, T=Z and p=3, then pR as 1+p=20, but ept0=2t which is negative for odd values of t. If however pR+, then eptt0>0, restoring the positivity property. Note that if T=R, then any function pR+ since 1+μtpt=1>0.

TDynamic Eq. (3)eptt0
Ry=ptyexpt0tpsds
ZΔy=ptyi=t0t11+pi
q0yΔ=ptyst0tT1+sq1ps

Table 5.

The exponential function for the continuous, discrete, and quantum time scale (q>1), assuming pR.

Some of the properties of the time scales exponential function are consistent with the convenient properties in the continuous case. If p,qR and t,sT, then

  1. e0ts=1, eptt=1,

  2. epqts=eptseqts,

  3. epts=epst=1epts,

  4. eptreprs=epts,

  5. epσts=1+μtptepts.

Theorem 3.1. [See [3, Theorem 2.39]] If pR and a,b,cT, then

abpteptcΔt=epbcepacabptepcσtΔt=epcaepcb.

As an application of linear, homogeneous, first order dynamic equations, one may consider the Malthusian growth model. In “An essay on the principle of population” from 1798, Thomas Robert Malthus proposed an exponential law of population growth with the corresponding differential equation

P=rP,Pt0=P0,

where P is the population at time t, r is the inherent growth rate, and P0 is the initial population level at time t0R. This linear, homogeneous, first order differential equation has the solution Pt=ertt0P0. Assuming a positive initial population level P0>0, it follows that for a positive growth rate r>0, the population increases exponentially. If instead r<0 and P0>0, then the population goes extinct as limtertt0P0=0. Despite its simplicity and the unrealistic behavior of unbounded population levels for r > 0, the Malthusian model can sometimes serve short-term predictions.

Let us now consider the corresponding time scales model (3) with initial condition Pt0=P0>0 and inherent growth rate r>0, that is, PΔ = rP with Pt0=P0 for t0T. The respective solution is then Pt=ertt0P0, which is unbounded for r,P0>0, see Figure 2. Thus, for r,P0>0, the behavior of the solution is consistent with the solution in the continuous case. However, for r<0, the population does not have to go extinct but can result in biologically unmeaningful behavior as solutions can become negative.

Figure 2.

The behavior of the solution to PΔ=rP with Pt0=P0 where r = 0.45, t0 = 1, and P0 = 0.1, for T=R, T=Z and T=1.30. The solid line represents the solution in the continuous case, the open circle represents the solution in the discrete case, and the stars represent the solution in the quantum calculus case with q=1.3.

Using the time scales exponential function that solves a linear, homogeneous, first order dynamic equation, we can use the variation of constants formula to obtain the solution to a linear, nonhomogeneous, first order dynamic equation.

Theorem 3.2. [See [3, Theorems 2.74 & 2.77]] Suppose pR, fCrd,t0T and y0R then the unique solution to

yΔ=pty+ft,yt0=y0

is given by

yt=eptt0y0+t0teptσsfsΔs.

Furthermore, the unique solution to

yΔ=ptyσ+ft,yt0=y0

is given by

yt=eptt0y0+t0teptsfsΔs.

Example 3.3. Suppose that the life span of a certain species is one time unit. Suppose that just before the species dies out, eggs are laid that are hatch after one time unit. The species is therefore only alive on T=k=02k2k+1, see also [3, Example 1.39] and [10]. Suppose further that during the specie’s, life cycle, the species reduces due to external factors with rate d ∈ (0, 1) and at the end of the life cycle t=2k+1, the individuals alive in (2k,2k+1) lay eggs that result in the reproduction rate r>0. The corresponding dynamic equation for the species N(t) at time t, is then

NΔt=ptNt,withpt=dt2k2k+1rt=2k+1

and initial condition N0=N0. We note that even though pt is discontinuous at t=2k+1, ptR. Theorem 3.2 gives the population at time t2m2m+1 as

Nt=N0eptt0=N0ept2mk=0m1ep2k+12kep2k+22k+1=N0exp2mtddsk=0m1exp2k2k+1dds1+r=N0edtm1+rm.

Example 3.4. Newton’s law of cooling suggests that the temperature of an object at time t, Tt, changes dependent on the temperature of its surrounding, Tm. Then, Tt=κTTm, where κ is the heat transfer coefficient. Suppose that an object with initial temperature T0 is cooled in a lab environment. Due to safety regulations, once the lab assistant leaves the work space, the object can only be exposed to an environment that preserves the current temperature of the object. The cooling of the object can be modeled using time scales with the underlying time domain to be the working hours of the lab assistant. Assume that the lab assistant’s working hours, and therefore the time scale, is of the form T=i=0aibicidi, where the interval aibi are the working hours prior to lunch, and cidi are the working hours of the lab assistant after lunch of day i. One way of modeling this scenario on time scales is

TΔ=ptTTm,pt=κtaibicidi0tbidi

with initial temperature Tt0=T0 for t0T. Since pt is rd-continuous and regressive, the theorems above can be applied despite the discontinuity of pt.

Example 3.5. The following example is from [11], where a Keynesian-Cross model with lagged income is considered. Here, the aggregated income y changes according to

yΔ=δdσty,tt0T,

where dt is the aggregated demand at time t and δ01 is the “adjustment speed”. Since dt can be expressed as the addition of aggregated consumption (c), aggregated investment (I), and governmental spending (G), we have dt=ct+I+G for I,G0. Under the assumption that aggregated consumption is itself linear in the aggregated income, we have ct=a+byt with a,b>0 so that the model reads as

yΔ=δa+byσ+I+Gy.

Under the assumption that pt:=1δbμt0, we can apply yσ=y+μyΔ, and express the dynamic equation as

yΔ=δa+I+Gpt+δb1pty.

which is a linear, non-homogeneous, first order dynamic equation. It is left as an exercise to apply the techniques of this subsection to derive an explicit solution to this dynamic equation.

Example 3.6. Let us consider a time scales analogue of the popular logistic growth model y=ry1yK, namely,

yΔ=ryσ1yK,yt0=y0,E5

with growth rate r > 0, and carrying capacity K > 0, and initial population size yt0>0 at time t0T. Even though this is an example of a nonlinear dynamic equation of first order, we can apply the substitution z=1y for y0, to obtain the linear dynamic equation

zΔ=yΔyyσ=rz+rK,zt0=1y0.

For rR, the solution is then given by Theorem 3.2. Using also Theorem 3.1 and resubstituting yields

yt=y0Kertt0Ky0+y0.E6

It can be easily checked that yt0=y0 and that y solves (5), see also [12].

Note that for T=R, (5) collapses to the Verhulst model y=ry1yK and the solution (6) reads in this case as

yt=y0Kertt0Ky0+y0,

which coincides with the classical solution.

3.2 Linear systems

Let us now consider (2) with f:T×Rn×RnRn for nN=1,2,3. In order to extend the solution methods for linear first order dynamic equations that were introduced in the previous section for scalar functions, the definitions of rd-continuity and delta differentiability have to be first extended to matrix valued functions A:TRm×n. This adjustment is mostly proposed element-wise. More precisely, A is rd-continuous on T if aij is rd-continuous on T for all 1in, 1jm. The class of all such rd-continuous m×n-matrix-valued functions on T is then denoted by CrdTRm×n. Similarly, we say that A is delta differentiable (or short: differentiable), if aij is delta differentiable for all 1in, 1jm. Similar to the scalar case, the following identity holds for any matrix-valued (delta) differentiable function A,

Aσt=At+μtAΔt.

The property of regressive is however not defined elementwise. Instead, we say that ARn×n is regressive if In+μtAt is invertible for all tTκ, where InRn×n is the identity matrix. The class of rd-continuous and regressive functions is denoted by RTRn×n (or short R).

Note that even if all entries of A are regressive, A does not have to be regressive. Take for example T=Z with

A=a11a12a21a22=0223.

Then all entries are regressive as 1+aij0 for all 1i,j2 but detI+A=0.

As for the scalar case, differentiation is linear, that is,

αA+βBΔt=αAΔt+βBΔt

for differentiable m×n-matrix-valued functions A,B, and α,βR.

We consider

yΔ=AtyE7

to be the system analogue of (3). If A is n×n matrix valued function, then, the unique solution to (7) with y(t0)=In, where In is the n×n identity matrix, is denoted by yt=eAtt0. If ARn×n and T=R then eAtt0=eAtt0, and if T=Z, then eAtt0=I+Att0. The analogue of (4) in higher dimensions is

yΔ=Atyσ,

where A(t) is the conjugate transpose of A(t).

Theorem 3.7. (See [3, Theorems 5.24 & 5.27]). Let ARTRn×nRn×n and suppose that f:TRn is rd-continuous. Let t0T and y0Rn. Then, the initial value problem

yΔ=Aty+ft,yt0=y0

is given by

yt=eAtt0y0+t0teAtστfτΔτ.

The unique solution to

yΔ=Atyσ+ft,yt0=y0

is given by

yt=eAtt0y0+t0teAtτfτΔτ.

Example 3.8. In [13], the authors consider the Cucker-Smale type model on an isolated T (i.e., every tT is isolated) with supT= and supμt:tT<,

xiΔ=viviΔ=1Nj=1Naijvjvi,E8

where aijR0+=0 and i12N represents the impact of agent’s j opinion onto the agent’s i opinion. The variable xi represents the state of agent i, and vi is the consensus parameter of agent i. The original Cucker-Smale model, see [14], is a discrete time system discussing the flock behavior of birds, where vi represents the velocity of bird i and xi is its position. The weights aij quantify the way the birds influence each other.

Note that since T is isolated, we can equivalently write (8) as

xiσt=xit+μtvit,viσt=vit+μtNj=1Naijvjtvit,

or in form of a system in y=x1x2xNv1v2vNT,

yΔ=By,B=1N0NNIN0NAD,,E9

where Aij=aij for i,j12N, D=diagd1d2dN with dk=j=1Nakj, 0N is a matrix of dimension N×N with all entries being zero, and IN is the identity matrix of dimension N×N.

If BR, then the solution to (9) with initial condition y(t0) = y0 is y(t) = eB (t,t0) y0. In order for BR, NIN + μ(t)(A-D) must be invertible because

B˜t=I2N+μtB=INμtIn0NCt,Ct=IN+μt1NAD,

and

detB˜t=detI2N+μtB=detINdetCt.

We conclude this section by examples of nonlinear dynamic equations that can be transformed into a system of linear dynamic equations of first order, so that Theorem 3.7 provides its solution.

Example 3.9. Let T be again an isolated time scale, that is, every point in T is isolated and infμt:tT>0. Consider

xσk=Kx1μtαK+μtαx,E10

with initial values x0=x0x1xk10k, K > 0, and –αR+. Eq. (10) is a delayed Beverton-Holt model and can be used to model mature individuals of a population, assuming that it takes k reproductive cycles for an individual to become mature, where the length of a reproductive cycle starting at t is μt. An application may be populations where the lengths between breeding cycles is temperature dependent. Model (10) has been considered in [15] (and, for T=Z, in [16]), where the authors applied the transformation yKx for x0 to obtain

YΔ=AtY+btwithAt=1μt0k1Ik1μαs,bt=0k1α,E11

where s=k1k2k3kk1 and 0k1Rk1×1 is vector of zeros. Applying Theorem 3.7, to (11) yields the solution.

Example 3.10. In [17], the authors proposed the following nonlinear system of dynamic equations to model the spread of a contagious disease,

SΔ=βtSσIνtS+γtI+νtκ,IΔ=βtSσIγtIνtI.

In line with well-established epidemic models, the population was compartmentalized into susceptible S and infected I individuals. The model assumes that the disease is spread by contact with an infected individual with a transmission rate of β>0. The recovery rate is assumed to be γ>0 and recovered individuals rejoin the group of susceptible individuals. The death rate is νt across the population and νtκ newborns join the group of susceptibles.

By introducing a new variable wS+I, wΔ=νtw+νtκ. This first order, linear, nonhomogeneous dynamic equation can be solved using Theorem 3.2, assuming νtR. The solution is then wt=eνtt0I0+S0κ+κ, so that, after recalling that S=wtI, the dynamic equation in I can be expressed as

IΔ=βtwσIσIγtIνtI.

Although the dimension has been reduced to one, the dynamic equation is still nonlinear. Defining however y=1I for I0 yields again a linear dynamic equation, namely

yΔ=βtwσt+γt+νtyσ+βt.

Applying Theorem 3.2 gives the solution

yt=eptt0y0+t0teptsβsΔs,

where pt=βtwσtγt+νt is assumed to be an element of R. Resubstituting yields then the solution I and using S=wI yields S.

For more epidemic models on time scales that are systems of first order nonlinear dynamic equations, see [18, 19, 20, 21]. While the dynamic Susceptible-Infected-Recovered epidemic model introduced in [18] can be solved explicitly via variable transformations, in most cases, including [19], explicit solutions to nonlinear dynamic equations are not available. In these cases, properties of solutions such as existence and uniqueness are of fundamental interest. The interested reader is referred to [22, Section 2] and [3, Section 8.2].

References

  1. 1. May R. Simple mathematical models with very complicated dynamics. Nature. 1976;261:459-467
  2. 2. Hilger S. Analysis on measure chains—A unified approach to continuous and discrete calculus. Research in Mathematics. 1990;18:18-56
  3. 3. Bohner M, Peterson A. Dynamic Equations on Time Scales. Boston, MA: Birkhäuser Boston Inc.; 2001 ISBN 0-8176-4225-0. An introduction with applications
  4. 4. Historical el nino/la nina episodes (1950–present), climate prediction center. Available from: https://origin.cpc.ncep.noaa.gov/products/analysis_monitoring/ensostuff/ONI_v5.php [Accessed: March 15, 2019]
  5. 5. Anderson DR, Krueger RJ, Peterson AC. Delay dynamic equations with stability. Advances in Difference Equations. 2006;2006:19
  6. 6. Kelley WG, Peterson AC. Difference Equations: An Introduction with Applications. Boston, MA: Academic Press, Inc.; 1991 ISBN 0-12-403325-3
  7. 7. Pötzsche C. Chain rule and invariance principle on measure chains. Journal of Computational and Applied Mathematics. 2002;141(1):249-254 ISSN 0377-0427. Dynamic Equations on Time Scales
  8. 8. Keller S. Asymptotisches Verhalten invarianter Faserbundel bei Diskretisierung und Mittelwertbildung im Rahmen der Analysis auf Zeitskalen. Augsburg: Universität Augsburg; 1999. Thesis (Ph.D.)
  9. 9. Agarwal RP, Bohner M. Basic calculus on time scales and some of its applications. Results in Mathematics. 1999;35(1–2):3-22
  10. 10. Christiansen FB, Fenchel TM. Theories of Populations in Biological Communities. Ecological Studies, Berlin Heidelberg: Springer; 2012. Available from: https://books.google.de/books?id=HAL8CAAAQBAJ
  11. 11. Tisdell CC, Zaidi A. Basic qualitative and quantitative results for solutions to nonlinear, dynamic equations on time scales with an application to economic modelling. Nonlinear Analysis. 2008;68(11):3504-3524
  12. 12. Bohner M, Warth H. The Beverton–Holt dynamic equation. Applicable Analysis. 2007;86(8):1007-1015
  13. 13. Girejko E, Machado L, Malinowska AB, Martins N. On consensus in the Cucker-Smale type model on isolated time scales. Discrete & Continuous Dynamical System Series. 2018;11(1):77-89. ISSN 1937-1632
  14. 14. Cucker F, Smale S. Emergent behavior in flocks. IEEE Transactions on Automatic Control. 2007;52(5):852-862. DOI: 10.1109/TAC.2007.895842
  15. 15. Bohner M, Cuchta T, Streipert S. Delay dynamic equations on isolated time scales and the relevance of one-periodic coefficients. Mathematicsl Methods in the Applied Sciences. 2022:1-18. DOI: 10.1002/mma.8141
  16. 16. Bohner M, Dannan FM, Streipert S. A nonautonomous Beverton-Holt equation of higher order. Journal of Mathematical Analysis and Applications. 2018;457(1):114-133
  17. 17. Bohner M, Streipert S. An integrable SIS model on time scales. In: Bohner M, Siegmund S, Šimon Hilscher R, Stehlík P, editors. Difference Equations and Discrete Dynamical Systems with Applications. Cham: Springer International Publishing; 2020. pp. 187-200
  18. 18. Bohner M, Streipert S, Torres DFM. Exact solution to a dynamic sir model. Nonlinear Analysis Hybrid Systems. 2019;32:228-238
  19. 19. Ferreira RAC, Silva CM. A nonautonomous epidemic model on time scales. Journal of Difference Equations and Applications. 2018;24(8):1295-1317
  20. 20. Sae-Jie W, Bunwong K, Moore E. The effect of time scales on sis epidemic model. WSEAS Transactions on Mathematics. 2010;9(10):757-767
  21. 21. Yeni G. Modeling of HIV, SIR and SIS Epidemics on Time Scales and Oscillation Theory. ProQuest LLC, Ann Arbor, MI: Missouri University of Science and Technology; 2019. ISBN 978-1392-67226-6. Thesis (Ph.D.)
  22. 22. Lakshmikantham V, Kaymakçalan B, Sivasundaram S. Dynamic Systems on Measure Chains, Volume 370 of Mathematics and its Applications. Dordrecht: Kluwer Academic Publishers; 1996

Written By

Sabrina Streipert

Reviewed: 25 March 2022 Published: 15 March 2023