sub-harmonic resonance and multi-annual oscillations in northern mammals: a non-linear...

14
Sub-harmonic resonance and multi-annual oscillations in northern mammals: a non-linear dynamical systems perspective q W.M. Schaer a,b, * , B.S. Pederson a , B.K. Moore c , O. Skarpaas d , A.A. King b , T.V. Bronnikova a a Department of Ecology and Evolutionary Biology, University of Arizona, Tucson, AZ 85721, USA b Program in Applied Mathematics, University of Arizona, Tucson, AZ 85721, USA c School of Education, University of Arizona, Tucson, AZ 85721, USA d Department of Mathematics, University of Arizona, Tucson, AZ 85721, USA Abstract We conjecture that the well-known oscillations (3- to 5-yr and 10-yr cycles) of northern mammals are examples of subharmonic resonance which obtains when ecological oscillators (predator-prey interactions) are subject to periodic forcing by the annual march of the seasons. The implications of this hypothesis are examined through analysis of a bare-bones, Hamiltonian model which, despite its simplicity, nonetheless exhibits the principal dynamical features of more realistic schemes. Specifically, we describe the genesis and destruction of resonant oscillations in response to variation in the intrinsic time scales of predator and prey. Our analysis suggests that cycle period should scale allometrically with body size, a fact first commented upon in the empirical literature some years ago. Our calculations further suggest that the dynamics of cyclic species should be phase coherent, i.e., that the intervals between successive maxima in the corresponding time series should be more nearly constant than their amplitude – a prediction which is also consistent with observation. We conclude by observing that complex dynamics in more realistic models can often be continued back to Hamiltonian limits of the sort here considered. Ó 2000 Elsevier Science Ltd. All rights reserved. 1. Introduction The multi-annual oscillations (Fig. 1) of northern mammals [1–7] are arguably among ecology’s most celebrated patterns. Recently, Norrdahl [8] proposed a general explanation: the cycles result from the in- teraction of specialist predators and their victims, with the intrinsic time scale of the prey species setting the period of oscillation. This hypothesis is noteworthy on two counts. In the first place, as Norrdahl explains, it oers a spartan accounting of the principal empirical findings. To a reasonable approximation, it further articulates the conclusions to which one is led by considering seasonal predator–prey interactions from the viewpoint of non-linear dynamics. In the present paper, we enunciate these conclusions and sketch their derivation. In this regard, we observe that the mathematical formulation of Norrdahl’s hypothesis is that northern mammal cycles are ecological examples of sub-harmonic resonance. From this principal conclusion follow two subsidiary results: that cycle period should scale allometrically with body size; and that the intervals between successive maxima in the time series of cyclic species should often be more nearly constant www.elsevier.nl/locate/chaos Chaos, Solitons and Fractals 12 (2001) 251–264 q This paper is dedicated to William A. Calder, III who gracefully responded to our criticism [26] of one of his papers [9], even though he was right all along and surely must have known it. * Corresponding author. E-mail address: [email protected] (W.M. Schaer). 0960-0779/00/$ - see front matter Ó 2000 Elsevier Science Ltd. All rights reserved. PII: S 0 9 6 0 - 0 7 7 9 ( 0 0 ) 0 0 0 6 2 - X

Upload: others

Post on 24-Apr-2021

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Sub-harmonic resonance and multi-annual oscillations in northern mammals: a non-linear ...blc.arizona.edu/courses/schaffer/Csf/Subharmonicres.pdf · 2016. 3. 11. · Sub-harmonic

Sub-harmonic resonance and multi-annual oscillations innorthern mammals: a non-linear dynamical systems

perspective q

W.M. Scha�er a,b,*, B.S. Pederson a, B.K. Moore c, O. Skarpaas d, A.A. King b,T.V. Bronnikova a

a Department of Ecology and Evolutionary Biology, University of Arizona, Tucson, AZ 85721, USAb Program in Applied Mathematics, University of Arizona, Tucson, AZ 85721, USA

c School of Education, University of Arizona, Tucson, AZ 85721, USAd Department of Mathematics, University of Arizona, Tucson, AZ 85721, USA

Abstract

We conjecture that the well-known oscillations (3- to 5-yr and 10-yr cycles) of northern mammals are examples of subharmonic

resonance which obtains when ecological oscillators (predator-prey interactions) are subject to periodic forcing by the annual march of

the seasons. The implications of this hypothesis are examined through analysis of a bare-bones, Hamiltonian model which, despite its

simplicity, nonetheless exhibits the principal dynamical features of more realistic schemes. Speci®cally, we describe the genesis and

destruction of resonant oscillations in response to variation in the intrinsic time scales of predator and prey. Our analysis suggests that

cycle period should scale allometrically with body size, a fact ®rst commented upon in the empirical literature some years ago. Our

calculations further suggest that the dynamics of cyclic species should be phase coherent, i.e., that the intervals between successive

maxima in the corresponding time series should be more nearly constant than their amplitude ± a prediction which is also consistent

with observation. We conclude by observing that complex dynamics in more realistic models can often be continued back to

Hamiltonian limits of the sort here considered. Ó 2000 Elsevier Science Ltd. All rights reserved.

1. Introduction

The multi-annual oscillations (Fig. 1) of northern mammals [1±7] are arguably among ecology's mostcelebrated patterns. Recently, Norrdahl [8] proposed a general explanation: the cycles result from the in-teraction of specialist predators and their victims, with the intrinsic time scale of the prey species setting theperiod of oscillation. This hypothesis is noteworthy on two counts. In the ®rst place, as Norrdahl explains,it o�ers a spartan accounting of the principal empirical ®ndings. To a reasonable approximation, it furtherarticulates the conclusions to which one is led by considering seasonal predator±prey interactions from theviewpoint of non-linear dynamics. In the present paper, we enunciate these conclusions and sketch theirderivation. In this regard, we observe that the mathematical formulation of Norrdahl's hypothesis is thatnorthern mammal cycles are ecological examples of sub-harmonic resonance. From this principal conclusionfollow two subsidiary results: that cycle period should scale allometrically with body size; and that theintervals between successive maxima in the time series of cyclic species should often be more nearly constant

www.elsevier.nl/locate/chaos

Chaos, Solitons and Fractals 12 (2001) 251±264

q This paper is dedicated to William A. Calder, III who gracefully responded to our criticism [26] of one of his papers [9], even though

he was right all along and surely must have known it.* Corresponding author.

E-mail address: [email protected] (W.M. Scha�er).

0960-0779/00/$ - see front matter Ó 2000 Elsevier Science Ltd. All rights reserved.

PII: S 0 9 6 0 - 0 7 7 9 ( 0 0 ) 0 0 0 6 2 - X

Page 2: Sub-harmonic resonance and multi-annual oscillations in northern mammals: a non-linear ...blc.arizona.edu/courses/schaffer/Csf/Subharmonicres.pdf · 2016. 3. 11. · Sub-harmonic

than their amplitude. As evidenced by Fig. 1 and the literature [9] both predictions are consistent withempirical observation.

2. Resonance

Resonance [10,11] is the phenomenon, familiar to any parent who ever pushed a child on a swing,whereby an oscillator (the swing) subject to periodic forcing (the pushing), vibrates at a frequency which isrationally related to that of the forcing. In the case of the child on the swing, the two frequencies are usuallythe same. This is called 1:1 or ``harmonic'' resonance, and it is the case most familiar to population bi-ologists [12]. Other possibilities include sub-harmonic resonance, in which instance, the frequency of theresonant vibrations is lower than that of the forcing, and superharmonic resonance, in which instance, theresonant frequency is greater.

In the case of northern mammal cycles, the presumptive oscillator is a predator±prey interaction, whileperiodic forcing is imposed by the annual march of the seasons. That seasonality is an essential componentto any mechanistic model of cyclic species follows from the animals' natural history. In the majority ofcases, reproduction is con®ned to the spring and summer with the consequence that population densitiesdecline during the winter. In short, there is ``phase-locking'' between the seasonal zeitgeber and the eco-logical oscillator. From this biologically pedestrian observation follow unexpected consequences.

3. A bare bones model

The essential phenomenology of seasonal predator±prey systems can be captured by straightforwardgeneralization of the well-known model of Volterra [13]. Following [14], we write 1

dx=dt � lx�1� � sin2pt ÿ ey�;dy=dt � ly�ex ÿ 1�; �1�

where x and y, are the natural logarithms of the densities of victims and predators. Of the three parameters,� is the magnitude of seasonality ± assumed to act on the victims' per capita rate of increase ± while lx andly , specify the two species' intrinsic time scales. These are the ``size-related'' rates of increase to whichNorrdahl refers. The derivation of these equations from a more traditional formulation of Volterra'sequations is described in the footnote 1.

Eq. (1) omit biologically important considerations such as predator satiation and victim self-limitation.These omissions notwithstanding, Eq. (1) nonetheless constitute the dynamical birthing place whereinoriginate the periodic behaviors of more realistic models [14,15]. The latter include a detailed representation[16] of snowshoe hare demography which can reproduce the cycle's ``natural'' period and amplitude as wellas its response to experimental manipulation [7]. In short, there are reasons to believe that analyzing Eq. (1)can contribute to our understanding of multi-annual oscillations in nature.

1 Derivation of Eq. (1): A more traditional formulation of Volterra's model may be written as follows:

dP=dT � P�bkV ÿ d�;dV =dT � V �r�T � ÿ kP�:

Here V and P are the densities of predators and victims, k is the per predator kill rate, b converts harvested victims into baby predators

and d and r(T) are (respectively), per capita rates of predator mortality and victim reproduction. To model the e�ects of seasonality, we

set r�T � � r0�1� � sin2pxT �, where x is the frequency, 1 yÿ1, of seasonal variations in victim reproduction and � indexes their

magnitude. Now de®ne dimensionless densities, v � V =V � and p � P=P ��0�, where V � � d=bk is the equilibrial density of victims, and

P ��0� � r=k is the corresponding density of predators when T � 0; 1; 2; . . .. Further de®ne dimensionless time, t � xT . Substituting

these quantities into Eq. (0) yields Eq. (1), where lx � �r0=x�;ly � �d=x� and t � xT .

252 W.M. Scha�er et al. / Chaos, Solitons and Fractals 12 (2001) 251±264

Page 3: Sub-harmonic resonance and multi-annual oscillations in northern mammals: a non-linear ...blc.arizona.edu/courses/schaffer/Csf/Subharmonicres.pdf · 2016. 3. 11. · Sub-harmonic

Following Norrdahl, we focus on the inherent time scales of the interacting species and consider thee�ects of lx and ly on the periods of resonant oscillations induced by Eq. (1). Without loss of generality, weintroduce new parameters, l and �l, such that

lx � l�1� ��;ly � l=�1� ��: �2�

Thus, l is the geometric mean of lx and ly , i.e., l � �lxly�1=2, while �l quanti®es their disparity. The utility

of this substitution will become apparent shortly.

4. The non-seasonal case

Ecologists have long been familiar [13,17] with non-seasonal versions of Eq. (1). In this case, � � 0, andthe x±y phase plane is densely populated by closed, neutrally stable solution curves surrounding the centerat the origin (Fig. 2). The period, T, of these cycles increases monotonically, and without bound, as onemoves outward from the origin. Close to the origin,

T � Tmin � 2p=l: �3�

5. Computing an orbit's rotation number

It is convenient to think of the solution curves in Fig. 2 as degenerate tori and to associate with each ofthem a rotation number, q. We do this by sampling solution curves at intervals equal to the period of theforcing which is imposed once � exceeds 0. Such a construction is called a ``stroboscopic map''. This ac-complished, we select an orbit and compute its rotation number as indicated in Fig. 3. Speci®cally, weassign an angle, hi to each point, i � 1; 2; 3; . . . on the orbit. Then q is calculated according to the formula

q � limN!1�1=N�

XN

i�0

�hi�1 ÿ hi�: �4�

Fig. 1. Historical time series for Canadian lynx (Lynx canadensis) (top) and grey-sided vole (Clethrionomys rufocanus) (bottom). Lynx

data (pelts shipped from the McKenzie River District by Hudson's Bay) from Elton and Nicholson [33]. Vole data (animals trapped

per 100 trap nights in Fennoscandia) from Hansen et al. [34].

W.M. Scha�er et al. / Chaos, Solitons and Fractals 12 (2001) 251±264 253

Page 4: Sub-harmonic resonance and multi-annual oscillations in northern mammals: a non-linear ...blc.arizona.edu/courses/schaffer/Csf/Subharmonicres.pdf · 2016. 3. 11. · Sub-harmonic

From a biological point of view, q can be viewed as the inverse of an orbit's ``average period'', i.e., the meaninterval between successive maxima in the associated time series. More fundamentally, the concept allowsus to divide the degenerate tori into classes according to whether q is rational or irrational. The signi®canceof this distinction, as we next discuss, is that it allows us to comprehend the consequences of introducingseasonality.

6. Consequences of seasonality

Implicit in much of the ecological literature is the assumption that the e�ect of seasonality is to replaceequilibrial dynamics with annual cycles which vary from year to year only to the extent that one year'sweather, and hence mean rates of reproduction and mortality, di�ers from the next. In fact, as shown inFig. 4a, the dynamical consequences of seasonality are far more complicated. In particular, we note thefollowing:1. For small values of �, the irrational tori are una�ected, save that they are no longer degenerate. Motion

on the irrational tori is quasi-periodic, i.e., the associated time series evidence oscillations modulated bya second, incommensurate frequency.

2. Whereas the irrational tori survive the imposition of in®nitesimal seasonality, their rational counterpartsare destroyed. In their stead, one observes complicated structures called ``island chains''. Each chain is

Fig. 2. Dynamics of Eq. (1) in the absence of seasonality, i.e., with � � 0. An in®nite number of neutrally stable cycles surround the

center at the origin.

Fig. 3. Calculating an orbit's rotation number. Each point, i, in the �x; y� time series is assigned an angle, hi, by surperposing the

data on coordinate axes as shown in the ®gure. The orbit's rotation number is the average change, �1=N�P�hi�1 ÿ hi�, in the limit of

large N.

Fig. 4. (a) Principal features of the non-integrable phase space viewed as a stroboscopic map. The map may be thought of as a slice of

a three-dimensional space, i.e., tori appear as closed loops; cycles as ®nite numbers of periodic points. Tori which survive the transition

to non-integrability are all irrational. Replacing the rational tori are island chains organized about neutrally stable and saddle cycles.

The former manifest themselves in section as elliptic ®xed points (eigenvalues of magnitude 1); the latter as hyperbolic points (ei-

genvalues real). In this picture, the stochastic layers, while still bounded by irrational tori, have grown to macroscopic proportions.

This re¯ects the continuing breakdown of irrational tori which attends increasing degrees of non-integrability. Note also, the existence

of secondary island chains which themselves are surrounded by tertiary chains, etc. (b). Predator±prey dynamics of Eq. (1) for weak

(� � 0:05) seasonal forcing. l � 1:0; �l � 0. (c) Predator±prey dynamics of Eq. (1) for strong (� � 1:0) seasonal forcing.

l � 1:0; �l � 0.

c

254 W.M. Scha�er et al. / Chaos, Solitons and Fractals 12 (2001) 251±264

Page 5: Sub-harmonic resonance and multi-annual oscillations in northern mammals: a non-linear ...blc.arizona.edu/courses/schaffer/Csf/Subharmonicres.pdf · 2016. 3. 11. · Sub-harmonic

W.M. Scha�er et al. / Chaos, Solitons and Fractals 12 (2001) 251±264 255

Page 6: Sub-harmonic resonance and multi-annual oscillations in northern mammals: a non-linear ...blc.arizona.edu/courses/schaffer/Csf/Subharmonicres.pdf · 2016. 3. 11. · Sub-harmonic

organized about a pair of periodic orbits, the rotation number of which equals that of the rational torusit replaces. One of these cycles is neutrally stable, while the other has the stability character of a saddle.

3. Associated with the saddle cycles are chaotic orbits. For small values of �, these orbits are bounded byirrational tori and consequently con®ned to thin, so-called ``stochastic'' layers.

4. Surrounding the neutrally stable cycles are secondary tori which themselves are surrounded by addition-al island chains. In fact, the island chains are self-similar, i.e., the primary island chains are surroundedby secondary chains which themselves are surrounded by tertiary chains, etc. This fact is a consequenceof the Poincar�e±Birkho� theorem [18].With increasing seasonality, the irrational tori also begin breaking down in a sequence determined by

their rotation numbers [18]. As a result, chaotic orbits, heretofore con®ned to thin layers, now wander morewidely, eventually merging to form regions of macroscopic chaos referred to as the ``stochastic sea''. Forsmall values of �, the sea is kept far from the origin by surviving irrational tori (Fig. 4b). For larger valuesof �, the stochastic sea laps closer to the origin (Fig. 4c). At the same time, the island chains also begin todisappear, with the outermost chains being the ®rst to su�er destruction.

These results exemplify the transition from ``integrable'' to ``non-integrable'' dynamics in Hamiltoniansystems as described by the KAM theorem [10,11,18,19] and, as such, they are very general. In such casesthey provide a realistic description of real-world phenomena. By contrast, the relevance of Hamiltoniandynamics to ecology depends on the fact [14±16] that a number of ecological models possess Hamiltonianlimits and that the dynamics to be found therein can be continued into biologically plausible regions ofparameter space. Fundamentally, the apparent ubiquity of Hamiltonian limits in ecology re¯ects two facts.The ®rst is that energy ¯ows up the trophic ladder; the second, that predator±prey interactions, at theircore, are essentially glori®ed pendula.

7. Progenitrix of cycles

In Section 6, we observed that an in®nite number of cycles and, hence, the island chains organized aboutthem, are created when � is varied away from zero by an arbitrarily small number. Of course, an in®nitenumber of cycles does not correspond to all possible cycles. Indeed, we already know from Eq. (3) that thecycles thus created have rotation numbers, q < l=2p < 1=Tmin. It follows that increasing l results inthe creation of additional cycles. Where do these cycles come from? The answer is surprisingly simple. All ofthe cycles emerge from the annual cycle which corresponds to the central ®xed point of the stroboscopicmap.

In the case of the period-2 island chain, the chain's origination is a two step process involving super- andsub-critical period-doublings that occur at di�erent values of l. Analogous complications attend the originsof the period-3 and period-4 chains [20±22]. The remaining sub-harmonics are created by single bifurca-tions. This process is illustrated in Fig. 5 wherein we show the birth of cycles of rotation numbersq � 1=5; 1=6 and 1=7. These are, of course, but three of an in®nite number of cycles which arise on the rangeof l values depicted. For example, between l � l1=5 (birth of the 1/5 cycle) and l1=6 (birth of the 1/6 cycle),

Fig. 5. Origination of multi-annual cycles in response increasing the geometric mean of the species intrinsic time scales. Here

� � 0:2; �l � 0.

256 W.M. Scha�er et al. / Chaos, Solitons and Fractals 12 (2001) 251±264

Page 7: Sub-harmonic resonance and multi-annual oscillations in northern mammals: a non-linear ...blc.arizona.edu/courses/schaffer/Csf/Subharmonicres.pdf · 2016. 3. 11. · Sub-harmonic

is l2=11 (birth of the 2/11 cycle); between l1=5 and l2=11, is l3=16, etc. This enumeration can be continued adin®nitum, with the consequence that on any interval, l � �l1; l2�, there arise an in®nite number of orbitswith periods ranging from 2p=l1 to 2p=l2. These orbits are born as imperceptible modulations of theannual cycle and thereafter increase in amplitude (Fig. 6).

In the course of this evolution, each cycle is destabilized by period-doubling, and there follows a period-doubling cascade (Fig. 7). The cascades accumulate, i.e., the interval of l values between successive dou-

Fig. 6. Successive stages in the evolution of the 12- and 13-year cycles in response to changing values of l. Elliptic cycles indicated by

circles; Saddle cycles by crosses. Top left: Birth of the 13 year cycle, l � 0:4835; top right: l � 0:4880; bottom left: Birth of the 12 year

cycle, l � 0:5240; bottom right: l � 0:5300.

Fig. 7. Period-doubling of the 1:5 resonant cycle in response to increasing the geometric mean of the species' intrinsic time scales, l.

Here � � 0:2; �l � 0.

Fig. 8. Origin and destabilization of the 1:5 resonance. The resonant Hopf (vertical line at left) and ®rst three period-doublings curves

(right) are shown.>Fig. 8. Origin and destabilization of the 1:5 resonance. The resonant Hopf (vertical line at left) and ®rst three

period-doublings curves (right) are shown.

W.M. Scha�er et al. / Chaos, Solitons and Fractals 12 (2001) 251±264 257

Page 8: Sub-harmonic resonance and multi-annual oscillations in northern mammals: a non-linear ...blc.arizona.edu/courses/schaffer/Csf/Subharmonicres.pdf · 2016. 3. 11. · Sub-harmonic

blings diminishes [23], so that for each value of q, there is a range of parameter values for which cycles ofthat rotation number are neutrally stable. As shown in Fig. 8, this range depends on both l and the degreeof seasonality, �.

These observations allow us to associate a range of available periods (Fig. 9) with a given l±� pair. Theshortest is Tmin � 2p=l; the longest corresponds to the accumulation point of the above-mentioned period-doubling cascade. To summarize: the e�ects of parametric variation are to down-shift the range of availableperiods in response to increasing l and to narrow the range in response to increasing �.

From a deterministic point of view, it is the destabilization of the cycles and their period-doubled de-scendants which sets the maximum value of l consistent with observable motions of a particular rotationnumber. In real-world systems, an equally important determining factor is that with increasing values of lminimum population densities become so small that the biological expectation is extinction. Happily, bothmechanisms work in the same direction.

8. Asymmetry in time scales

Thus far, we have assumed that the intrinsic time scales of the two species are the same. Energeticconsiderations dictate that this will often be roughly true, i.e., as a generality, really large predators do notsubsist on really small prey. There are, however, exceptions, in which instances, the parameter �l in Eq. (2)will di�er from zero by a signi®cant amount. The consequence of this complication is to narrow the range ofavailable periods ± principally by moving the period-doubling curves to the left.

9. Mathematical basis of resonance and self-similarity

Mathematically, the birth of sub-harmonics can be understood in terms of the eigenvalues (also calledFloquet multipliers) of the annual cycle when viewed stroboscopically. At l � 0, both eigenvalues equal 1.Thereafter, they separate and move to the left along the unit circle in the complex plane. With each value ofl,one can therefore associate an angle, hl. With increasing l; hl increases from 0 to p, thereby passingthrough an in®nite number of values rationally related to 2p. At each such value, a pair of cycles is emittedfrom the central periodic orbit. Eventually, the eigenvalues remerge at ÿ1, at which point the period-2 sub-harmonic is born. This explains the origination of the primary island chains as shown in Fig. 5.

If we now consider the eigenvalues of the cycles about which the primary island chains are organized, wesee that the foregoing process is repeated. With increasing values of l, the eigenvalues of the various sub-harmonics move along the unit circle, again passing through an in®nite number of values rationally relatedto 2p. At these points, secondary island chains are produced. By the same mechanism, the secondary sub-harmonics give rise to tertiary island chains, which, in turn give rise to fourth order chains, etc. This ex-plains the self-similarity of the island chains. Each of the period-doubled cycles also emits its own set ofsecondary island chains, which then give rise to higher-order chains. In short, a single mechanism gives riseto all possible sub-harmonics.

Fig. 9. A range of periods [Tmin, Tmax] is associated with every �l; �� pair. Tmin � 2p=l; Tmax corresponds to the point at which the

period-doubling sequence accumulates.

258 W.M. Scha�er et al. / Chaos, Solitons and Fractals 12 (2001) 251±264

Page 9: Sub-harmonic resonance and multi-annual oscillations in northern mammals: a non-linear ...blc.arizona.edu/courses/schaffer/Csf/Subharmonicres.pdf · 2016. 3. 11. · Sub-harmonic

10. Allometric scaling of cycle period

Norrdahl's view that cycle period is determined by the intrinsic time scales echoes a suggestion by Calder[9] that mammalian herbivores oscillate at periods which scale allometrically with body size. Calder ob-served that, if cycle period, T, is measured in years and body mass, M, in kg,

T � aMb; �5�where a ranges 6.5±7.7 and b from 0.25 to 0.27. This result is based on data for 10 species, several of whichevidence multi-annual cycles. 2

Our results suggest a simple explanation for this ®nding, subject to the important proviso that thepopulations in question participate in exploitative interactions. In this regard, we note that, for mostspecies, the intrinsic per capita rate of increase, r, varies with body mass, M, according to

r � cMÿd �6�[24]. For homeotherms, when r is measured in units of yÿ1 and M in kg, c � 2:2 and d � 0:275. This allowsus to obtain a lower bound on cycle period by computing Tmin. If we follow Norrdahl and equate l with thevictims' per capita rate of increase, rv, we obtain

Tmin � 2p=rv � 2p=cMÿd � 2:9M0:275; �7�which, on a log±log plot, gives the proper slope, but the wrong intercept. 3 Despite their utter simplicity,Eq. (1) thus give order of magnitude agreement with the empirical results.

11. From Hamiltonian to dissipative dynamics

What happens when Eq. (1) are perturbed in the direction of greater biological realism by addition of a®nite victim carrying capacity, Kv, and predator satiation? This question was answered by King andScha�er [15] who considered the system

dx=dt � lx�g�t� ÿ aex ÿ �1ÿ a�f �x�ey �;

dy=dt � ly �f �x�ex ÿ 1�; �8�

f �x� � �1� g�=�1� gex�;

g�t� � 1� � sin2pt:

Here, a is the ratio of V�, the equilibrium density of victims under predation in a constant environment, toKv, and g determines the rate at which per predator kill rate saturates with increasing numbers of victims.

If a � g � 0, Eq. (8) collapse to (1), for which reason, we say that a � g � 0 is a Hamiltonian limit, H. Atthis point, areas of the Poincar�e map are preserved, and there exist the in®nite numbers of regular andchaotic motions discussed above. For all other values of a and g, the so-called dissipative regime, areas ofthe Poincar�e map either expand or contract. As a consequence, neutrally stable limit sets are no longerpossible. Instead, one observes sinks, saddles and repellers. More speci®cally, the range of asymptotic statesdepends on the relative values of a and g. On varying these quantities, one observes the following:

2 Subsequent attempts [25] to apply Eq. (5) to wider range of taxa were methodologically ¯awed ± most of the populations for which

cycle periods were reported evidenced little objective evidence of periodic behavior [26].3 By this accounting, lemmings should cycle with periods on the order of a year, and the snowshoe hare, periods on the order of 3

years. These estimates are low by a factor of 3. They can be improved if we observe that predators are often somewhat larger than their

victims and replace rv with (lx ly�1=2.

W.M. Scha�er et al. / Chaos, Solitons and Fractals 12 (2001) 251±264 259

Page 10: Sub-harmonic resonance and multi-annual oscillations in northern mammals: a non-linear ...blc.arizona.edu/courses/schaffer/Csf/Subharmonicres.pdf · 2016. 3. 11. · Sub-harmonic

1. If a=g is su�ciently large, all initial states tend to the annual cycle which, on the Poincar�e map, is a ®xedpoint near the origin.

2. For smaller values of a=g, both the annual cycle and one or more sub-harmonic oscillations are stable.3. For still smaller values of a=g, the annual cycle is unstable. Trajectories based at points in its vicinity

spiral out to one or more attractors which may either be periodic or chaotic.4. As a=g is subjected to further reductions, the attractors, be they periodic or chaotic, move ever outward

from the origin, with the consequence that the average ``period'' of oscillation, i.e., the mean intervalbetween successive peaks in the associated time series, increases.These possibilities are illustrated in Fig. 10 which depicts the consequences of decreasing a while holding

g ®xed. In Fig. 10a, the ®xed point near the origin and the 1:3 cycle are stable; in Fig. 10b, the ®xed pointhas lost stability and initial conditions are attracted either to the 1:3 or to the 4:12 cycle which surrounds it.In Fig. 10c the stable cycles include 1:4, 3:12 and the 2:12 resonances. There is also a chaotic saddle [27]which can induce extended chaotic transients. Finally, in Fig. 10d, all of the cycles are unstable and motionis on a strange attractor.

12. Phase coherent character of multi-annual oscillations

Multi-annual cycles in nature are often characterized by the relative constancy of their period vis-a-vistheir amplitude (Fig. 1) an observation to which attention has been called in the case of lynx cycle byScha�er [28]. This property, called ``phase coherence'' in the dynamical systems literature, 4 is characteristicof the type of chaos ®rst described by Otto R�ossler [29]. Chaotic motions induced by Eq. (8) also possessthis property [15] for the reason that the rotation numbers of the periodic orbits about which they are

Fig. 10. Periodic and chaotic attractors in the dissipative regime are contained within a disk or annulus centered about the ®xed point

near the origin. With increasing values of the ratio, g=a, both the inner and outer regions of this region move outward. Top left: log

a � ÿ1:2; top right: log a � ÿ1:3; bottom left: log a � ÿ1:4; bottom right: log a � ÿ1:4332. In all cases, log g � ÿ1:1726.

4 The name re¯ects the fact that nearby initial conditions on a phase coherent attractor preserve their relative phase relations over

extended periods even if the motion is chaotic. This results in a power spectrum with sharp peaks superposed on a noisy background

[30].

260 W.M. Scha�er et al. / Chaos, Solitons and Fractals 12 (2001) 251±264

Page 11: Sub-harmonic resonance and multi-annual oscillations in northern mammals: a non-linear ...blc.arizona.edu/courses/schaffer/Csf/Subharmonicres.pdf · 2016. 3. 11. · Sub-harmonic

organized decrease monotonically as one moves outward from the annual cycle. 5 This geometry traces tothe Hamiltonian limit, H and is therefore to be expected in systems which can be formulated as pertur-bations of such a limit. It is not, we emphasize, a characteristic of chaotic dynamics generally.

The topology responsible for phase coherence in chaotic solutions to Eq. (8) is worth displaying, and wedo so in Fig. 11. Here, recurrent points are color-coded by the reciprocals of their rotation numbers. Moreprecisely, we display points which return to a small neighborhood of themselves after 5±18 iterations of thePoincar�e map. Color-coding is according to the sequence of colors in the visible spectrum: dark red cor-responds to points for which �1=q� � smin < 5; dark violet, to points for which �1=q� � smax > 9. Twofeatures of the diagram merit comment. The ®rst is that there is an overall tendency for the color to changefrom red to violet as one moves outward from the attractor's inner margin. This re¯ects the aforementionedordering of the saddle cycles. The second is that swirls of red spread out from the inner margin while swirlsof violet spread inward from the outer margin. To understand why this is the case, we recall that the cyclesabout which the attractor is organized have the stability character of saddles, with the consequence thatrecurrent points will be on or near the stable manifold of such a cycle. What our construction thus reveals isthe tangled structure of the manifolds, which is the geometric basis of chaotic behavior. From the viewpoint

Fig. 11. Recurrent points of the Poincar�e map color-coded by the reciprocals of their rotation numbers. Revealed are the homoclinic

tangles which are the topological basis of chaotic motion.

5 Chaotic attractors are organized about in®nite numbers of saddle periodic orbits in the sense that the latter are ``dense on the

attractor.'' By this it is meant that every point on such an attractor is arbitrarily close to a periodic orbit.

W.M. Scha�er et al. / Chaos, Solitons and Fractals 12 (2001) 251±264 261

Page 12: Sub-harmonic resonance and multi-annual oscillations in northern mammals: a non-linear ...blc.arizona.edu/courses/schaffer/Csf/Subharmonicres.pdf · 2016. 3. 11. · Sub-harmonic

of the biology, what is important, of course, is not the tangling, but the fact that motion on the resultingattractor is an amalgamation of nearly periodic motions representing a circumscribed range of frequencies.

13. Relation of the present model to three-trophic level models of the lynx±hare cycle

Thus far, our discussion has focused on predator±prey interactions. In fact, recent opinion regarding thegenesis of multi-annual cycling in nature has tended to emphasize the importance of resource limitation aswell as predation. This perspective is most strongly supported by ®eld experiments of Krebs and his as-sociates [7] on snowshoe hare (Lepus americanus) populations in the Canadian Yukon. Principally, theseauthors showed that supplemental feeding of hares dramatically increased their maximum density andprolonged the peak phase of the cycle. At the same time, excluding mammalian, but not avian, predators,slowed the post-peak decline. In other words, the experiments suggest a three-trophic-level system. Sta-tistical analysis [31] of historical data points to the same conclusion: hare population cycles appear to bedetermined by two factors whereas explication of lynx time series requires but a single factor. This resulthas been interpreted to indicate that the hare is limited both by its food supply and by predation, lynxpopulation growth rates depend only on the abundance of the hare which is its principal food species.Elsewhere [16], we have shown that a three trophic level model of snowshoe hare demography is, in fact,compatible not only with the available census data, but also with the experimental results noted above. Thepresent analysis, of course, makes no mention of herbivore forage and it is therefore worthwhile, we believe,to enquire as to what relationship the dynamics of Eq. (1) might bear to those of a more expansive scheme.

Interestingly, the nature of this relationship turns out to hinge on the dynamics of the forage. In the caseof the snowshoe hare, the limiting resource is winter browse consisting of the terminal twigs of woodyplants [5]. Essentially, there are two qualitatively distinct ways in which one can model browse dynamics.On the one hand, one can suppose that the consumption of browse has a minimal e�ect on the plants'photosynthetic abilities. This leads to an equation of the form

dB=dt � rB�t��KB ÿ B� ÿ FB�H ; t�B: �9�Here B and H are the densities of browse and hares, rB(t) is the seasonally-dependent rate of browseproduction, KB, the maximum abundance of browse and FB(), the hare's functional response which is as-sumed to depend on both hare abundance and on the time of year.

An alternative approach is to assume that browse consumption has a signi®cant e�ect on the plants'photosynthetic abilities. This leads to a browse renewal equation of the form

dB=dt � rB�t�B��1ÿ B=KB� ÿ FB�H ; t�B�: �10�The di�erences between (9) and (10), while they would appear slight, have important implications. Withbrowse renewing according to Eq. (9), the hare and its food supply cannot oscillate in the absence ofpredators. As a result, the oscillations of the three level system trace to a Hamiltonian limit which corre-sponds to Eq. (1). In this fundamental sense, the oscillations are predator±prey cycles, and it is interesting,therefore, that the model of King and Scha�er, which seems capable of giving a good accounting of theempirical observations, caricatures browse dynamics according to Eq. (9).

In contrast, the use of Eq. (10) makes for the possibility of multiple Hamiltonian limits [15]. One of theseis an herbivore±vegetation system; another, a predator±prey interaction. In this case, explicating the dy-namics of the three level model may require consideration of both limits and therefore all three statevariables. In passing, we remark that the possibility of multiple Hamiltonian limits raises fascinating, and,as yet unexplored, questions regarding the interaction of invariant motions emanating therefrom.

14. Ubiquity of Hamiltonian limits in ecological models

Our approach to the dissipative dynamics of Eq. (8) is motivated by the fact that non-integrableHamiltonian limits can be found in a variety of ecological models. We emphasize this point because it is so

262 W.M. Scha�er et al. / Chaos, Solitons and Fractals 12 (2001) 251±264

Page 13: Sub-harmonic resonance and multi-annual oscillations in northern mammals: a non-linear ...blc.arizona.edu/courses/schaffer/Csf/Subharmonicres.pdf · 2016. 3. 11. · Sub-harmonic

completely at variance with the conventional wisdom [17,32]. The latter holds that Hamiltonian modelsby virtue of their structural instability [11] are irrelevant to the behavior of non-Hamiltonian schemeswhich attempt to a more faithful representation of biological reality. In fact, structural stability is amathematical concept which is less relevant to the dynamics of non-linear systems than is often sup-posed [15].

In addition to Eq. (8), non-integrable Hamiltonian limits have been observed in the following instances:1. Periodically forced chemostat (double Monod) models in which holozoic protozoan predators consume

heterotrophic bacteria.2. Periodically forced three-level food chains in which the vegetation renews logistically [15].3. The snowshoe hare model mentioned above [16].4. N P 2 species systems consisting of two or more predator±prey pairs coupled by overlapping predator

diets in non-seasonal environments.The ®nal example merits comment in that it is autonomous (constant parameter values) and therefore

illustrates the general principle [10] that coupling two non-linear oscillators is fundamentally equivalent tosubjecting a single oscillator to periodic forcing. An example ± two predators, two prey ± is shown in Fig. 1wherein we display what is called a two-dimensional ``surface of section'' of the full four-variable system.To our knowledge, Fig. 12 is the ®rst observation of the ®eldmarks of non-integrable Hamiltonian dy-namics in ecological models. It suggests that there may be merit in revisiting the statistical mechanicalmodels of Kerner and Leigh (see [17] for review) of n-species ecological associations from a modern dy-namical point of view.

In conclusion, we note that the apparent ubiquity of Hamiltonian limits in ecological models re¯ects thefact that embedded in predator±prey interactions are the non-integrable dynamics of the frictionless pen-dulum. It is a surprising result, suggesting as it does an unsuspected unity among disparate scienti®c dis-ciplines.

Acknowledgements

This work was supported by a grant from the National Science Foundation to WMS and by FlinnFoundation Fellowships in Biomathematics to AAK.

Fig. 12. Non-integrable dynamics in an autonomous system (constant environment) consisting of two predator±prey pairs cou-

pled by overlapping predator diets. The same mix of regular and chaotic motion observed in seasonal predator±prey models is ob-

served.

W.M. Scha�er et al. / Chaos, Solitons and Fractals 12 (2001) 251±264 263

Page 14: Sub-harmonic resonance and multi-annual oscillations in northern mammals: a non-linear ...blc.arizona.edu/courses/schaffer/Csf/Subharmonicres.pdf · 2016. 3. 11. · Sub-harmonic

References

[1] Elton C. Voles, mice and lemmings. Oxford: Oxford University Press; 1942.

[2] Keith LB. Wildlife's 10 year cycle. Madison: University of Wisconsin Press; 1963.

[3] Finerty JP. The population ecology of cycles in small mammals. New Haven, CT: Yale University Press; 1980.

[4] Keith LB. Role of food in hare population cycles. Oikos 1983;40:385±95.

[5] Keith LB. Dynamics of snowshoe hare populations. In: Current mammalogy, vol. 2. New York: Plenum Press, 1990. p. 119±195.

[6] Hanski I, Korpim�aki E. Microtine rodent dynamics in northern Europe parameterized models for the predator±prey interaction.

Ecology 1995;76:840±50.

[7] Krebs CJ, Boutin S, Boonstra R, Sinclair ARE, Smith JNM, Dale MRT, Tukington R. Impact of food and predation on the

snowshoe hare cycle. Science 1995;269:1112±5.

[8] Norrdahl K. Population cycles in northern mammals. Biol Rev Camb Philos Soc 1995;70:621±37.

[9] Calder WA. An allometric approach to population cycles of mammals. J Theoret Biol 1983;100:275±82.

[10] Lichtenberg AJ, Lieberman MA. Regular and chaotic dynamics. New York: Springer; 1992.

[11] Guckenheimer J, Holmes P. Nonlinear oscillations dynamical systems and bifurcations of vector ®elds. New York: Springer; 1983.

[12] Nisbet RM, Gurney WSC. Population dynamics in a periodically varying environment. J Theoret Biol 1976;56:59±475.

[13] Volterra V. Fluctuations in abundance of a species considered mathematically. Nature 1926;118:558±60.

[14] King AA, Scha�er WM, Treat J, Gordon C, Kot M. Weakly dissipative predator±prey systems. Bull Math Biol 1996;58:835±60.

[15] King AA, Scha�er WM. The rainbow bridge: Hamiltonian limits and resonance in predator±prey dynamics. J Math Biol

1999;39:439±69.

[16] King AA, Scha�er WM. Geometry of a population cycle: a mechanistic model of snowshoe hare demography. Ecology (in press).

[17] MaynardSmith J. Models in ecology. London: Cambridge University Press; 1974.

[18] H�enon M. Numerical explorations of Hamiltonian systems. In: Iooss G, Helleman RHG, Stora R, editors. Chaotic behavior of

deterministic systems. Amsterdam: North-Holland, 1983. p. 53±170.

[19] Tabor M. Chaos and integrability in nonlinear systems. New York: Wiley; 1989.

[20] Gambaudo JM. Perturbation of a Hopf bifurcation by an external time-periodic forcing. J Di� Eqs 1985;57:172±99.

[21] Vance WN, Ross J. Bifurcation structures of periodically forced oscillators. Chaos 1991;1:445±53.

[22] Kuznetsov YA. Elements of applied bifurcation theory. New York: Springer; 1995.

[23] Helleman RHG. Self-generated chaotic behavior in nonlinear systems. In: Cohen EGD, editor. Fundamental problems in

statistical mechanics, vol. 5. Amsterdam: North-Holland, 1980. p. 165±233.

[24] Fenchel T. Intrinsic rate of natural increase the relationship with body size. Oecologia 1974;14:317±26.

[25] Peterson RO, Page RE, Dodge KM. Wolves moose and the allometry of population cycles. Science 1984;224:1350±2.

[26] Krukonis G, Scha�er WM. Population cycles in mammals and birds does periodicity scale with body size. J Theoret Biol

1991;148:469±93.

[27] Kantz H, Grassberger P. Repellers semi-attractors and long-lived chaotic transients. Physica D 1985;17:75±86.

[28] Scha�er WM. Order and chaos in ecological systems. Ecology 1985;66:93±106.

[29] R�ossler OE. An equation for continuous chaos. Phys Lett 1976;35a:397±8.

[30] Farmer D, Crutch®eld J, Froehling H, Packard N, Shaw R. Power spectra and mixing properties of strange attractors. Ann New

York Acad Sci 1980;357:453±72.

[31] Stenseth N-Chr, Falck W, Bjùrnstad ON, Krebs CJ. Population regulation in snowshoe hare and canadian lynx asymmetric food

web con®gurations between hare and lynx. Proc Natl Acad Sci USA 1997;94:5147±52.

[32] May RM. Complexity and stability in model ecosystems. Princeton: Princeton University Press; 1973.

[33] Elton CS, Nicholson M. The 10-year cycle in numbers of the lynx in Canada. J Anim Ecol 1942;11:215±44.

[34] Hansen TF, Stenseth N-Chr, Henttonen H. Multiannual vole cycles and population regulation during long winters an analysis of

seasonal density dependence. Am Nat 1999;154:129±39.

264 W.M. Scha�er et al. / Chaos, Solitons and Fractals 12 (2001) 251±264