Pdf-Version - Centre de Recherches sur la Cognition Animale

Transcription

Pdf-Version - Centre de Recherches sur la Cognition Animale
Institut national agronomique Paris-Grignon
THÈSE
pour obtenir le grade de
Docteur de l’Institut national agronomique Paris-Grignon
présentée et soutenue publiquement par
Christian Jost
le 11 décembre 1998
Comparaison qualitative et quantitative de modèles
proie-prédateur à des données chronologiques en
écologie
Comparing predator-prey models qualitatively and
quantitatively with ecological time-series data
Jury :
Roger Arditi (professeur INA-PG, directeur de thèse) . . . . . . . . . . . . . . . . . .
Jean Clobert (directeur de recherches CNRS) . . . . . . . . . . . . . . . . . . . . . . . . . .
Donald L. DeAngelis (professeur USGS, University of Miami, rapporteur)
Pierre-Henri Gouyon (professeur Université Paris XI, président) . . . . . . . .
Claude Lobry (professeur Université de Nice) . . . . . . . . . . . . . . . . . . . . . . . . . .
François Rodolphe (directeur de recherches INRA, rapporteur) . . . . . . . . .
Éric Walter (directeur de recherches CNRS) . . . . . . . . . . . . . . . . . . . . . . . . . . . .
♥
“Che voust far, cunter il vent nu poust pischar”
Old Rhaetian saying
“Steter Tropfen höhlt den Stein”
Old German saying
Acknowledgements
Although there appear only two names besides mine (Roger Arditi and Ovide Arino)
in this thesis, many more people have contributed to this work during the last 32 years
and 10 months. Since this place might be my only opportunity to thank them in an
official document with worldwide distribution I beg the readers comprehension for this
lengthy list. Additional thanks go to:
The people in Dietingen (47 35’ 30” N, 8 49’ E) for a great childhood.
The teachers in primary school that organized inspiring skiing camps and thus
directed my professional career towards the educational sector. Unfortunately
there are no skiing camps in higher education (sigh!).
Peter Zimmermann for teaching great natural sciences.
Herbert Amann and Pierre Gabriel for teaching me mathematical thinking.
Anthony Burgess and J. R. R. Tolkien for their tremendeous writings that frequently prevented me from working during weekends.
The laboratory in Orsay, especially the secretaries and surface technicians, for
keeping the infrastructure running and assisting my integration into the French
system.
Roger Arditi for being a thoughtful and patient thesis adviser even during his
double activities in Lausanne and Paris.
The participants of the spring schools in Luminy (CNRS-GDR 1107) for listening
to my beginners problems under the spring sun of Southern France.
The members of the jury (see cover page) to take the time and effort to read
and comment on this work, especially François Rodolphe for …nding sense in my
rudimentary statistical knowledge and Don DeAngelis for taking the time to come
from Florida and to analyze critically my concepts.
The ETHICS library system and its staff of the polytechnical school in Zürich that
allowed extensive bibliogaphic research in the shortest possible time (including the
journey from Paris to Zürich).
The Swiss National Science Foundation to grant the resources for food, housing
and cinema during these studies.
4
Identifying predator-prey models (PhD Thesis)
C. Jost
My family with all its members (especially Mueti, Papo and Mamo) and my two
godchildren Rina and Demian for preventing my total immersion in dull scienti…c
questions.
And, for all the rest, Sergine . . . especially for persuading me with insistence that
protozoans do not believe in differential equations (the bets are still open).
Contents
I
General introduction and main results of this thesis
1 Predator dependence in the functional response and its implications
in ecology
1.1
1.2
2.2
2.3
3
Some comments on mathematical population ecology . . . . . . . . . . .
3
1.1.1
A general predator-prey modelling framework . . . . . . . . . . .
5
1.1.2
Top-down and bottom-up control in predator-prey models . . . .
6
Introducing predator dependence . . . . . . . . . . . . . . . . . . . . . .
7
1.2.1
A novel approach: ratio-dependent functional responses . . . . .
8
1.2.2
Experimental evidence for ratio dependence . . . . . . . . . . . .
10
1.2.3
Mechanistic derivations of ratio dependence . . . . . . . . . . . .
11
1.2.4
Alternatives to predator dependence . . . . . . . . . . . . . . . .
12
1.2.5
Ratio dependence: state of the art . . . . . . . . . . . . . . . . .
13
1.2.6
A short digression: to what density does ‘density dependence’
refer in the functional response? . . . . . . . . . . . . . . . . . .
15
2 A case of model selection
2.1
1
17
The candidate predator-prey models . . . . . . . . . . . . . . . . . . . .
17
2.1.1
18
Note on discrete versus continuous models . . . . . . . . . . . . .
Qualitative comparison of models
. . . . . . . . . . . . . . . . . . . . .
19
2.2.1
Interpretation in the predator-prey context . . . . . . . . . . . .
19
2.2.2
Adding a trophic level . . . . . . . . . . . . . . . . . . . . . . . .
20
2.2.3
Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
22
Boundary dynamics: do they matter? . . . . . . . . . . . . . . . . . . .
23
2.3.1
Some other models that offer deterministic extinction . . . . . .
24
2.3.2
Is there a biological control paradox? . . . . . . . . . . . . . . . .
25
2.4
Microbiologists did it
2.5
Quantitative comparison of models
2.5.1
. . . . . . . . . . . . . . . . . . . . . . . . . . . .
27
. . . . . . . . . . . . . . . . . . . .
30
A remark on model selection criteria . . . . . . . . . . . . . . . .
32
i
ii
Identifying predator-prey models (PhD Thesis)
C. Jost
2.5.2
The errors that govern our modelling world . . . . . . . . . . . .
32
2.5.3
Implementation of …tting algorithms . . . . . . . . . . . . . . . .
34
2.5.4
An ‘in silico’ approach to model selection . . . . . . . . . . . . .
35
2.5.5
The ‘in vivo’ data analysis . . . . . . . . . . . . . . . . . . . . . .
37
3 Concluding remarks and perspectives
If
3.1
General conclusions
3.2
Perspectives for continuation of work
43
. . . . . . . . . . . . . . . . . . . . . . . . . . . . .
43
. . . . . . . . . . . . . . . . . . .
44
3.2.1
A non-parametric approach . . . . . . . . . . . . . . . . . . . . .
44
3.2.2
Paying more attention to deterministic extinction . . . . . . . . .
45
3.2.3
Modelling plankton data more realistically . . . . . . . . . . . . .
45
3.2.4
How to treat process and observation error together?
46
. . . . . .
Introduction générale et résultats principaux
1f
1f Prédateur-dépendance de la réponse fonctionnelle et implications en
écologie
3f
1.1f Quelques commentaires concernant les modèles mathématiques en écologie des populations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3f
1.1.1f Un cadre général pour la modélisation de systèmes proie-prédateur 6f
1.1.2f Le contrôle de haut en bas et de bas en haut dans les modèles
proie-prédateur . . . . . . . . . . . . . . . . . . . . . . . . . . . .
7f
1.2f Introduction de la prédateur-dépendance . . . . . . . . . . . . . . . . . .
8f
1.2.1f Une approche nouvelle : la réponse fonctionnelle ratio-dépendante
9f
1.2.2f Données expérimentales en faveur de la ratio-dépendance . . . .
10f
1.2.3f Dérivations mécanistes de la ratio-dépendance
. . . . . . . . . .
12f
1.2.4f Des alternatives pour la prédateur-dépendance . . . . . . . . . .
13f
1.2.5f Ratio-dépendance : l’état de l’art
. . . . . . . . . . . . . . . . .
14f
1.2.6f Une petite digression : à quoi se refère le terme de “densité” dans
un réponse fonctionnelle densité-dépendante? . . . . . . . . . . .
16f
2f Un cas de sélection de modèle
19f
2.1f Les modèles proie-prédateur potentiels . . . . . . . . . . . . . . . . . . .
19f
2.1.1f Petite remarque : modèle discret et modèle continu . . . . . . . .
20f
2.2f Comparaison qualitative des modèles
. . . . . . . . . . . . . . . . . . .
21f
2.2.1f Interprétation dans le context proie-prédateur . . . . . . . . . . .
21f
2.2.2f Ajout d’un niveau trophique
23f
. . . . . . . . . . . . . . . . . . . .
C. Jost
iii
Identifying predator-prey models (PhD Thesis)
2.2.3f Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.3f Dynamiques de bord : quelle importance ?
. . . . . . . . . . . . . . . .
2.3.1f Quelques autres modèles offrant l’extinction
25f
. . . . . . . . . . .
27f
2.3.2f Y-a-t-il un paradoxe du contrôle biologique ? . . . . . . . . . . .
27f
2.4f Les microbiologistes l’ont fait . . . . . . . . . . . . . . . . . . . . . . . .
31f
2.5f Comparaison quantitative des modèles
. . . . . . . . . . . . . . . . . .
33f
2.5.1f Une remarque sur le critère de sélection de modèle . . . . . . . .
36f
2.5.2f Les erreurs qui gouvernent notre monde modélisateur . . . . . .
36f
2.5.3f Implémentation des algorithmes d’ajustement . . . . . . . . . . .
38f
2.5.4f Une approche ‘in silico’ de la sélection de modèle
. . . . . . . .
39f
2.5.5f L’analyse ‘in vivo’ . . . . . . . . . . . . . . . . . . . . . . . . . .
43f
3f Conclusions générales et perspectives
3.1f Conclusions générales
II
25f
47f
. . . . . . . . . . . . . . . . . . . . . . . . . . . .
47f
3.2f Perspectives pour la continuation des études . . . . . . . . . . . . . . . .
48f
3.2.1f Modéliser l’extinction comme un phénomène déterministe . . . .
49f
3.2.2f Modélisation plus réaliste du plancton . . . . . . . . . . . . . . .
49f
3.2.3f Comment tenir compte à la fois d’une erreur de processus et d’une
erreur d’observation? . . . . . . . . . . . . . . . . . . . . . . . .
50f
Detailed studies (accepted or submitted articles)
49
4 The clear water phase in lakes: a non-equilibrium application of alternative phytoplankton-zooplankton models
51
4.1
4.2
4.3
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
53
4.1.1
Seasonality of plankton in lakes . . . . . . . . . . . . . . . . . . .
53
4.1.2
Simple mathematical models of predator-prey interactions . . . .
55
Problem formulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
56
4.2.1
The alternative models . . . . . . . . . . . . . . . . . . . . . . . .
56
4.2.2
The required patterns . . . . . . . . . . . . . . . . . . . . . . . .
57
Model analysis and results . . . . . . . . . . . . . . . . . . . . . . . . . .
58
4.3.1
Dimensionless forms . . . . . . . . . . . . . . . . . . . . . . . . .
58
4.3.2
Isoclines . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
58
4.3.3
Equilibria . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
59
4.3.4
Stability of the non-trivial equilibrium . . . . . . . . . . . . . . .
59
4.3.5
Comparison of dynamic properties in the two models . . . . . . .
60
iv
Identifying predator-prey models (PhD Thesis)
C. Jost
4.3.6
Model trajectories and plankton seasonality . . . . . . . . . . . .
62
4.3.7
Seasonal changes of parameter values . . . . . . . . . . . . . . . .
63
Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
64
Appendix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
66
4.A Detailed matrix analysis
. . . . . . . . . . . . . . . . . . . . . . . . . .
66
4.A.1 Prey-dependent model . . . . . . . . . . . . . . . . . . . . . . . .
67
4.A.2 Ratio-dependent model . . . . . . . . . . . . . . . . . . . . . . .
67
4.B Some effects of a density dependent mortality rate . . . . . . . . . . . .
69
4.4
5 About deterministic extinction in ratio-dependent predator-prey models
71
5.1
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
73
5.2
The model and its equilibria . . . . . . . . . . . . . . . . . . . . . . . . .
74
5.3
Stability of the equilibria
. . . . . . . . . . . . . . . . . . . . . . . . . .
75
5.4
(0, 0) as an attractor . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
79
5.5
Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
81
6 Predator-prey theory: why ecologists should talk more with microbiologists
85
6.1
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
87
6.2
A short historical perspective . . . . . . . . . . . . . . . . . . . . . . . .
88
6.3
Lessons for ecology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
90
7 Identifying predator-prey processes from time-series
95
7.1
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
97
7.2
The alternative models . . . . . . . . . . . . . . . . . . . . . . . . . . . .
99
7.3
Materials and methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100
7.4
7.5
7.3.1
Arti…cial time-series . . . . . . . . . . . . . . . . . . . . . . . . . 100
7.3.2
Error functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
7.3.3
Model selection . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104
7.3.4
Parameter estimation . . . . . . . . . . . . . . . . . . . . . . . . 105
7.3.5
Algorithmic details . . . . . . . . . . . . . . . . . . . . . . . . . . 105
Analysis and results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107
7.4.1
Model identi…cation . . . . . . . . . . . . . . . . . . . . . . . . . 107
7.4.2
Parameter estimation . . . . . . . . . . . . . . . . . . . . . . . . 108
Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
Appendix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113
C. Jost
Identifying predator-prey models (PhD Thesis)
v
7.A Calculating the derivatives of the state variables with respect to the parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113
8 From pattern to process: identifying predator-prey models from timeseries data
115
8.1
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117
8.2
The alternative models
8.3
Materials and methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121
8.4
8.5
. . . . . . . . . . . . . . . . . . . . . . . . . . . 120
8.3.1
Time-series . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121
8.3.2
Error functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123
8.3.3
Simulation study, numerical methods and bootstrapping . . . . . 125
8.3.4
Model comparisons . . . . . . . . . . . . . . . . . . . . . . . . . . 126
Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 128
8.4.1
Protozoan data (simple batch cultures)
8.4.2
Arthropod data (spatially complex laboratory systems) . . . . . 129
8.4.3
Plankton data (complex lake plankton systems) . . . . . . . . . . 130
Discussion
. . . . . . . . . . . . . . 128
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130
Appendix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138
8.A Residual bootstrapping and IEP E
. . . . . . . . . . . . . . . . . . . . 138
A Collection of predator-prey models
141
A.1 The prey growth function . . . . . . . . . . . . . . . . . . . . . . . . . . 141
A.2 The functional response . . . . . . . . . . . . . . . . . . . . . . . . . . . 142
A.3 The numerical response . . . . . . . . . . . . . . . . . . . . . . . . . . . 146
A.4 Predator mortality . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 146
B Data from Lake Geneva
147
C Distinguishability and identi…ability of the studied models
157
C.1 Distinguishability
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 157
C.2 Identi…ability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 157
C.2.1 Predator-prey model with prey-dependent functional response . . 158
C.2.2 Predator-prey model with ratio-dependent functional response . 159
D Transient behavior of general 3-level trophic chains
161
D.1 The method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 161
D.2 The 3-level trophic chains and the PEG-model . . . . . . . . . . . . . . 162
D.2.1 Analysis of the prey-dependent model . . . . . . . . . . . . . . . 163
vi
Identifying predator-prey models (PhD Thesis)
C. Jost
D.2.2 Analysis of the ratio-dependent model . . . . . . . . . . . . . . . 164
List of Figures
1.1
Predictions in chained vessel experiments . . . . . . . . . . . . . . . . .
11
1.2
In‡uence of m on predator isocline . . . . . . . . . . . . . . . . . . . . .
14
2.1
3-level food chain showing PEG-model dynamics . . . . . . . . . . . . .
22
2.2
Food web with two prey types showing PEG-model dynamics . . . . . .
22
2.3
Types of per capita growth rates . . . . . . . . . . . . . . . . . . . . . .
25
2.4
Biological control with the ratio-dependent model . . . . . . . . . . . . .
27
2.5
Problems with non-likelihood regression . . . . . . . . . . . . . . . . . .
36
2.6
Quality of parameter estimates . . . . . . . . . . . . . . . . . . . . . . .
38
3.1
Expectation of SODE’s
. . . . . . . . . . . . . . . . . . . . . . . . . . .
47
1f.1 Prédictions des expériments en compartiments enchaînés . . . . . . . . .
12f
1f.2 In‡uence de m sur l’isocline du prédateur . . . . . . . . . . . . . . . . .
15f
2f.1 Dynamique du modèle PEG, chaîne trophique . . . . . . . . . . . . . . .
24f
2f.2 Dynamique du modèle PEG, deux types de proie . . . . . . . . . . . . .
25f
2f.3 Types du taux de variation de la croissance . . . . . . . . . . . . . . . .
28f
2f.4 Contrôle biologique avec le modèle ratio-dépendent . . . . . . . . . . . .
30f
2f.5 Problèmes avec la régression hors du concept de vraisemblance . . . . .
41f
2f.6 Qualité des paramètres estimés . . . . . . . . . . . . . . . . . . . . . . .
42f
3f.1 Espérance d’une EDSO
. . . . . . . . . . . . . . . . . . . . . . . . . . .
51f
4.1
PEG-model dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . .
54
4.2
General isoclines . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
58
4.3
Parameter regions for stable spiral sinks . . . . . . . . . . . . . . . . . .
61
4.4
Reaction to enrichment
. . . . . . . . . . . . . . . . . . . . . . . . . . .
62
4.5
Typical trajectories . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
63
4.6
Trajectories with seasonal predator death rate . . . . . . . . . . . . . . .
64
4.7
Typical observed dynamics in Lake Geneva . . . . . . . . . . . . . . . .
70
vii
viii
Identifying predator-prey models (PhD Thesis)
C. Jost
5.1
Isoclines of a ratio-dependent model . . . . . . . . . . . . . . . . . . . .
76
5.2
Origin as a saddle point . . . . . . . . . . . . . . . . . . . . . . . . . . .
81
5.3
Origin and non-trivial equilibrium attractive . . . . . . . . . . . . . . . .
82
5.4
Origin attractive, limit cycle . . . . . . . . . . . . . . . . . . . . . . . . .
82
5.5
Origin globally attractive . . . . . . . . . . . . . . . . . . . . . . . . . .
83
6.1
D. E. Contois . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
85
6.2
Isoclines in chemostats and predator-prey systems . . . . . . . . . . . .
93
7.1
Good …ts of wrong model . . . . . . . . . . . . . . . . . . . . . . . . . . 100
7.2
Types of …tting dynamic models to data . . . . . . . . . . . . . . . . . . 102
7.3
Identi…ability of arti…cial data: Summary . . . . . . . . . . . . . . . . . 108
7.4
Sensitivity analysis of error functions . . . . . . . . . . . . . . . . . . . . 109
7.5
Estimation of local stability . . . . . . . . . . . . . . . . . . . . . . . . . 110
8.1
Observation-error-…t and process-error-…t . . . . . . . . . . . . . . . . . 124
8.2
Dynamic states of the predator-prey systems . . . . . . . . . . . . . . . 127
8.3
Examples of …ts to data . . . . . . . . . . . . . . . . . . . . . . . . . . . 133
B.1 Plankton dynamics year 1986 . . . . . . . . . . . . . . . . . . . . . . . . 148
B.2 Plankton dynamics year 1987 . . . . . . . . . . . . . . . . . . . . . . . . 149
B.3 Plankton dynamics year 1988 . . . . . . . . . . . . . . . . . . . . . . . . 150
B.4 Plankton dynamics year 1989 . . . . . . . . . . . . . . . . . . . . . . . . 151
B.5 Plankton dynamics year 1990 . . . . . . . . . . . . . . . . . . . . . . . . 152
B.6 Plankton dynamics year 1991 . . . . . . . . . . . . . . . . . . . . . . . . 153
B.7 Plankton dynamics year 1992 . . . . . . . . . . . . . . . . . . . . . . . . 154
B.8 Plankton dynamics year 1993 . . . . . . . . . . . . . . . . . . . . . . . . 155
D.1 Possible transitions between extrema . . . . . . . . . . . . . . . . . . . . 163
D.2 Full transition graph of extrema . . . . . . . . . . . . . . . . . . . . . . . 164
List of Tables
1.1
Equilibrium predictions with enrichment . . . . . . . . . . . . . . . . . .
9
1f.1 Prédictions des équilibres avec enrichissement . . . . . . . . . . . . . . .
10f
4.1
Seasonal parameter trends . . . . . . . . . . . . . . . . . . . . . . . . . .
64
6.1
Same models in microbiology and ecology . . . . . . . . . . . . . . . . .
88
6.2
Studies using Contois’ model . . . . . . . . . . . . . . . . . . . . . . . .
92
7.1
CV ’s of …tted parameters . . . . . . . . . . . . . . . . . . . . . . . . . . 110
8.1
Sources of real data
8.2
Summary of model selection results . . . . . . . . . . . . . . . . . . . . . 128
8.3
Comparative studies in microbiology . . . . . . . . . . . . . . . . . . . . 135
8.4
Detailed model selection results . . . . . . . . . . . . . . . . . . . . . . . 136
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122
A.1 Prey growth functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143
A.2 Prey-dependent functional responses . . . . . . . . . . . . . . . . . . . . 144
A.3 Predator-dependent functional responses . . . . . . . . . . . . . . . . . . 145
A.4 Predator mortality functions
. . . . . . . . . . . . . . . . . . . . . . . . 146
ix
x
Identifying predator-prey models (PhD Thesis)
C. Jost
Part I
General introduction and main
results of this thesis
1
Chapter 1
Predator dependence in the
functional response and its
implications in ecology
1.1
Some comments on mathematical population ecology
One of the …rst mathematical descriptions of population dynamics can be found by
the end of the 18th century when Malthus (1798) introduced what is known today
as Malthusian growth: a population increases exponentially as long as resources are
unlimited. De…ning N (t) to be the population abundance at time t this growth can be
described by the differential equation
dN (t)
= rN (t)
dt
with growth rate r. But what happens if resources (e.g., space, food, essential nutrients) are limited? That limited resources can stop population growth was introduced
empirically by Verhulst (1838) in what is called today logistic growth,
dN
= rN (1
dt
N
).
K
(1.1)
The carrying capacity K de…nes some population abundance limit beyond which the
population growth rate becomes negative, while below it, this growth rate is positive.
Thus, K is some equilibrium value towards which the population abundance tends
to converge. K can also be interpreted as a measure of the available resources, the
logistic growth is therefore a model with donor control (Pimm 1982): the population is
controlled by its resources.
There is another way how a population can be prevented from growing exponentially:
there can be another population that consumes the …rst population at a rate exceeding
this …rst populations’ growth rate. Such a predator-prey interaction has …rst been
described by two persons working independently, Lotka (1924) and Volterra (1926).
3
4
Identifying predator-prey models (PhD Thesis)
C. Jost
With P (t) being the abundance of this second population, the predator, they described
the interaction by the set of differential equations
dN
dt
dP
dt
= rN
aNP
= eaNP
P
(1.2)
with predator attack rate a, conversion efficiency e (percentage of consumed prey biomass that is converted into predator biomass) and predator mortality rate . Depending on the initial abundances of prey and predator they will cycle eternally, passing
periodically through the initial values (neutral stability). The discovery of a cycling
mathematical predator-prey system coincided with Elton’s (1924) article on ‡uctuating
animal populations. This initiated the hypotheses that such ‡uctuations are caused by
predator-prey interactions. Hundreds of papers have investigated this hypotheses and
still much is unknown (see review in Chitty 1996).
However, in this thesis I am not interested in cycling populations. Recall that in the
logistic population growth model control is exercised from the bottom (donor control
or bottom-up). Consider now the prey equilibrium in the Lotka-Volterra predator-prey
system,
N? =
.
ea
Not only is the predator preventing its prey from growing in…nitely, its parameters , e
and a also de…ne the equilibrium state of the prey. This relation expresses mathematically what is often called top-down control.
There are no unique de…nitions in ecology (speci…cally in the context of trophic
chains) for the terms top-down control/effect and bottom-up control/effect. In general,
top-down control is used in the sense of Hairston et al. (1960), that is that every trophic
level has the potential to control/repress its prey to a low level that is independent of the
prey’s resources, except if it is itself controlled by its own predator (the ‘why the world
is green hypotheses’ in the case of three-level trophic chains, stating that herbivores
do not graze plants down to low levels, thus creating a brown world, because they are
controlled by carnivores). A generalisation of this to food chains of arbitrary length is
the exploitation ecosystem hypotheses (EEH, Fretwell 1977; Oksanen et al. 1981) where
trophic levels an odd number below the top trophic level are top-down controlled, while
the others are bottom-up controlled. A special case are cascading effects (Carpenter
et al. 1985; Carpenter & Kitchell 1994) that study how changes in the top trophic level
cascade several levels down the trophic chain. Bottom-up effects are the driving factor
in donor-controlled food chains (Pimm 1982; Berryman et al. 1995) where every trophic
level is uniquely controlled by its resources and its mortality rate is independent of its
predators density. For further information see the discussions in Hunter & Price (1992),
Strong (1992), Power (1992) and Menge (1992) or the well written synthesis in chapter II
of Ponsard (1998). In this thesis, I will talk of bottom-up control when the equilibrium
abundance of a trophic level is positively correlated with the resources on the …rst trophic
level in the studied system (e.g., increasing K in the case of logistic growth). Top-down
control of a population refers to the case when a change in its predators abundance or
C. Jost
Identifying predator-prey models (PhD Thesis)
5
parameters induces a change in the prey population equilibrium abundance. As will be
seen later, these de…nitions are not mutually exclusive, a population can be controlled
both top-down and bottom-up.
The two views, top-down and bottom-up, are of general interest because they make
very different predictions how trophic level equilibrium abundances change with increasing resources (see Table 1.1). For example, to reduce algal density in a eutrophic
lake with four trophic levels (algae, herbivorous zooplankton, carnivorous zooplankton,
planktivorous …sh), the bottom-up control proposes that the unique way to do so is to
reduce nutrient levels in the lake, while the top-down approach by Oksanen et al. (1981)
suggests that, on the contrary, fertilizing the lake will reduce algal density by increasing
herbivorous zooplankton density.
Coming now back to the Lotka-Volterra system: are all predator-prey systems of
the top-down control type indicated by this system? Or do there also exist predatorprey systems that correspond better to bottom-up or mixed types of control? These
questions are of a general interest, because all direct interactions between species at
different trophic levels of a food web are of a predatory nature. The particular choice of
the mathematical form describing the predator-prey interaction in such general food web
models can have profound impacts on the predictions of these models to perturbations
at the bottom (changing nutrient status) or the top (introduction or removal of top
predators), see preceding paragraph. Such predictions are used when deciding about
management policy of exploited natural populations or in conservation biology issues.
There is still few known about the respective importance of top-down or bottom-up
forces in natural systems (Menge 1992; Power 1992). In this situation, an ecologist
should try to make predictions that are independent of the type of control, or at least,
he should identify the predictions that are sensitive to the dominance of one of the forces.
In the following, I will introduce a predator-prey model with bottom-up characteristics
that can serve as an alternative to Lotka-Volterra types of predation. By confronting
this model and Lotka-Volterra type models qualitatively and quantitatively with timeseries data from observed predator-prey systems, I hope to gain some information about
the controlling mechanisms in these systems.
1.1.1
A general predator-prey modelling framework
The predator-prey models that will be studied in this thesis follow two general principles.
The …rst one is that population dynamics can be decomposed into birth and death
processes,
dN
= growth death.
dt
The second one is the conservation of mass principle (Ginzburg 1998), stating that
predators can grow only as a function of what they have eaten. With these two principles
we can write the canonical form of a predator-prey system as
dN
= f (N )N g(N, P )P N (N )N
dt
dP
(1.3)
= eg (N, P )P P (P )P
dt
6
Identifying predator-prey models (PhD Thesis)
C. Jost
with prey per capita growth rate f (N ), functional response g(N, P ) [prey eaten per
predator per unit of time, Solomon 1949] and natural mortalities N (N ) and P (P ).
The numerical response (predator per capita growth rate as a function of consumption)
is considered in this framework to be proportional to the functional response. Note that
the system (1.3) becomes the Lotka-Volterra system (1.2) when f (N ) = r, g(N, P ) =
aN , N (N ) = 0 and P (P ) = . Many other forms have been proposed in the literature
for the functions f , g and x (x ∈ {N, P }). See Appendix A for an (incomplete)
collection of functions I found in the literature. In particular the functional response g
has incited creative works, see Bastin & Dochain (1990) who list over 50 models.
Usually one considers consumption to be the major death cause for the prey. In
this case N (N ) can be neglected and set to 0 (as long as the predator exists). I will
follow this usage in the present thesis, replace the growth rate f (N ) by the standard
logistic growth (Verhulst 1838, see Figure 2.3) and also consider predator mortality to
be constant as in the Lotka-Volterra equations, giving the equations
dN
N
= r 1
N g(N, P )P
dt
K
dP
= eg (N, P )P P.
(1.4)
dt
1.1.2
Top-down and bottom-up control in predator-prey models
Traditionally, the functional response in system (1.4) is assumed to be a function of
prey abundance only, g(N, P ) = g(N ), e.g.
g(N ) = aN
aN
g(N ) =
1 + ahN
(Lotka 1924 )
(Holling 1959b )
(1.5)
(1.6)
(see Table A.2 for further examples). I will call this case a prey-dependent functional
response (Arditi & Ginzburg 1989). In this case, the prey equilibrium density is of the
form
N? = g 1
.
e
Thus, as in the Lotka-Volterra case, it is entirely de…ned by the predator’s parameters.
Graphically, this is expressed in the vertical predator isocline in Figure 4.2a. Increasing
carrying capacity K (enrichment) will only change the prey isocline and result in a
destabilization of the equilibrium, the so-called paradox of enrichment (Rosenzweig
1971). The traditional prey-dependent approach in the context of framework (1.4)
will therefore always result in a top-down controlled predator-prey model that becomes
destabilized under enrichment.
The previous analysis indicates where we have to change our system (1.4) to break
this pattern: the functional response must also depend on predator abundance, g =
g(N, P ). I will call this case predator dependence. Usually, higher predator density
C. Jost
Identifying predator-prey models (PhD Thesis)
7
leads to more frequent encounters between predators. This can cause a loss in predating
efficiency either simply by time lost to detect that the encountered organism is not
a prey (this time is thus not spent searching for other prey), or because of active
interference between predators (Beddington 1975; Hassell & Varley 1969). Therefore,
predator abundance should in‡uence the functional response negatively,
dg (N, P )
< 0.
dP
How this modi…cation changes the top-down pattern will be discussed in more detail
in the next section. There is a large literature on predator dependence of the functional response in both vertebrates and invertebrates (reviews in Hassell 1978a and in
Sutherland 1996).
In behavioural ecology one …nds also the inverse effect, predator facilitation. Cooperation between predators can pay off in increased reproductive success. Examples
are lions hunting in pairs rather than alone (Krebs & Davies 1993, p. 278) or patellid
limpets capturing drifting kelp and seaweed debris (Bustamante et al. 1995). However,
this thesis will only consider detrimental predator dependence, which seems to be more
frequent.
1.2
Introducing predator dependence
Predator dependence in the functional response has been introduced on many occasions.
The …rst one that had some echo in the literature was by Hassell & Varley (1969), who
proposed that the attack rate a should decrease with increasing predator density. They
proposed an exponential decrease with rate m,
a = P
m
.
(1.7)
Parameter m can be interpreted as an interference coefficient. The stronger predators
interfere one with each other, the larger m should be. While used at …rst only to explain
attack rates measured in the …eld, incorporating this attack rate into system (1.4) also
showed a stabilising effect (Beddington 1975; Beddington et al. 1978). There are two
technical problems with this formulation. First, there is a minor dimensional problem
(Beddington et al. 1978), which can be remedied by introducing a ‘dummy’ parameter
uP with dimension [P ] (representing one ‘unit’ predator) and rewriting (1.7) as
a=
P
uP
m
.
Usually, this ‘dummy’ parameter is not mentioned but tacitly assumed (Hassell 1978a,
p. 84). The second problem is more grave and concerns the mathematical analysis of
systems using this type of attack rate: due to the exponent m it is often impossible to
…nd the equilibria explicitly, thus rendering stability or other types of analyses much
more tedious (requiring the use of implicit function theory or other more sophisticated
techniques).
8
Identifying predator-prey models (PhD Thesis)
C. Jost
Both problems can be avoided by introducing predator dependence in the way proposed by Beddington (1975) and DeAngelis et al. (1975),
g(N, P ) =
aN
1 + ahN + cP
(1.8)
with prey handling time h (the same as in the usual Holling type II functional response
(1.6)) and an empirical constant c. Beddington (1975) derived this model based on
behavioural mechanisms, assuming that the predator spends its time either searching
for prey, handling prey or handling other predators (recognition, maybe interference).
c is in this context the product of predator encounter rate and predator handling time.
The system (1.4) together with this functional response can be analysed in the usual
way, solving for equilibria and testing stability (e.g., with the Routh-Hurwitz criteria, Edelstein-Keshet 1988). Doing this, DeAngelis et al. (1975) noted that in case
of 1 + ahN cP this system becomes donor controlled. Therefore, the functional response (1.8) provides a simple way to get bottom-up features in a predator-prey system.
However, it has rarely been used either in mathematical studies or in applications to
real ecological systems. Likely reasons are that the parameter c is difficult to estimate
and the mathematical analysis is not as simple as for example with Lotka-Volterra or
Holling type II functional responses because it is a function of two separate arguments,
N and P , forcing the use of partial differentiation.
1.2.1
A novel approach: ratio-dependent functional responses
A simpler way to include predator dependence has been proposed by Arditi & Ginzburg
(1989), modelling the functional response as a function of one argument only (as in the
traditional way), but this argument being now the ratio prey per predator,
N
.
(1.9)
g(N, P ) = g
P
This is a special case of predator dependence, and models using this type of functional
response are usually called ratio-dependent models (Arditi & Ginzburg 1989). g(x) is
in general a continuous, increasing and bounded function of its argument x, just as
in prey-dependent models. For example, a frequently used ratio-dependent functional
response is based on an analogy with the Holling type II model (1.6),
N
N/P
.
g
(1.10)
=
P
1 + hN/P
The simplicity of the approach allowed Arditi & Ginzburg (1989) to analyse general
food chains and to compare how the equilibria of the different trophic levels change with
changing primary productivity F := f (N )N . The results are summarised in Table 1.1.
It can be seen that in ratio-dependent models equilibria of all trophic levels increase with
higher productivity (independently of the food chain length), while prey-dependent food
chains show an alternating pattern with increase, no change and decrease. In particular,
the second highest trophic level is always top-down controlled, while trophic levels that
C. Jost
Identifying predator-prey models (PhD Thesis)
9
Table 1.1: Trends of equilibria of basal prey (N ? ) and predatory trophic levels
(P1? , P2? , etc.) with increasing primary productivity F in either prey-dependent
or ratio-dependent food chains. Arrows indicate the direction of change. Table
adapted from Arditi & Ginzburg (1989).
Prey-dependence
Ratio-dependence
Trophic level
2
3 4
5
2 3
4
5
?
P4
%
%
?
P3
% →
% %
P2?
% → %
% % %
P1?
% → % &
% % % %
N?
→ % & %
% % % %
are an even number below the top level are bottom-up controlled. The prey-dependent
pattern has been observed in a three-level microbial laboratory food chain (Kaunzinger
& Morin 1998), although a slight (unexplained) increase of the second trophic level with
enrichment was observed. Oksanen et al. (1981), who analysed prey-dependent food
chains in general, thought to …nd the pattern predicted by prey-dependent theory in
data of ecosystems of arctic-subarctic plant communities (Oksanen 1983). In the …eld,
the best data are available from freshwater lakes along a gradient of nutrient content.
Extensive studies have shown that the biomasses of all trophic levels (phytoplankton,
herbivorous zooplankton, carnivorous zooplankton, planktivorous …sh) increase along
a gradient of enrichment (McCauley et al. 1988; Mazumder 1994; Mazumder & Lean
1994; McCarthy et al. 1995; Harris 1996; Sarnelle et al. 1998). Thus, the observations
conform qualitatively with the predictions of the ratio-dependent model. A number
of explanations have been advanced to explain this pattern (McCauley et al. 1988;
Abrams 1993), predator dependence in the functional response is just one possibility.
However, ratio dependence offers one of the most parsimonious ways to model the
observed patterns. Note that the use of a ratio-dependent functional response in system
(1.4) also in‡uences the link between stability and enrichment (see chapter 5 for details):
a change in carrying capacity K only in‡uences the amplitude of trajectories, but not
the stability behaviour of the equilibrium. This signi…es that the paradox of enrichment
is closely linked to the vertical predator isocline and any variation from this will make
other responses than destabilization possible.
Modelling food webs rather than food chains is mathematically more complex.
Michalski & Arditi (1995a) and Arditi & Michalski (1995) developed mathematical
food webs with Holling type II, DeAngelis-Beddington and ratio-dependent functional
responses. With the latter a distinctive feature emerged that was not present with the
other functional responses: while the dynamics are in a transient state (before reaching
equilibrium), trophic links appeared and disappeared frequently (Michalski & Arditi
1995b). Once at equilibrium, there are few links in the food web, but if the food web
is observed continuously during transient dynamics, the links accumulate and give very
complex looking food webs such as the cod food web in Lavigne (1996) (reproduction
in Yodzis (1995)).
10
1.2.2
Identifying predator-prey models (PhD Thesis)
C. Jost
Experimental evidence for ratio dependence
Besides the cited plankton abundances that are correlated with the trophic state of lakes
there is also experimental evidence. There has been a series of experiments with two
cladoceran predator species, Daphnia magna and Simocephalus vetulus, feeding on the
algae Chlorella vulgaris (Arditi et al. 1991b; Arditi & Saïah 1992). Algae were raised in
a separate vessel and added at a constant rate to a …rst vessel with the predator. Out‡ow
from this vessel (with algae only, predators prevented from migration to the next vessel)
went to a second vessel with predators and so on. The hypotheses was that in case of a
vertical predator isocline (prey dependence) the predator should subsist only in the …rst
vessel, grazing algae down to a …xed level that cannot sustain predator populations in
the following vessels. On the other side, if the predator isocline is slanted, predators are
likely to survive also in consecutive vessels (see Figure 1.1). The predators chosen for
the experiment had a particular spatial behaviour, Daphnia swimming homogeneously
in the medium while Simocephalus keeps close to the vessel walls. Daphnia followed the
pattern predicted by a vertical predator isocline, while Simocephalus persisted in several
consecutive vessels. In a next step Daphnia were prevented from moving in the whole
vessel while Simocephalus were distributed homogeneously by frequent stirring. With
this setup, Daphnia persisted in several vessels, while Simocephalus only persisted in the
…rst one. These experiments clearly demonstrate that spatial heterogeneity can cause
a slanted predator isocline. However, they do not show that the predator isocline goes
through the origin, as predicted by ratio-dependent models (Ruxton & Gurney 1992;
Holmgren et al. 1996). Figure 1.1 illustrates this point for the example of three consecutive vessels. Both the ratio-dependent model and a DeAngelis-Beddington type model
predict qualitatively the same pattern, namely that predator abundance at equilibrium
decreases geometrically from one vessel to the next.
In another study, Arditi & Akçakaya (1990) reanalyzed data of functional response
experiments found in the literature (predator-prey and host-parasitoid systems). In all
these experiments both prey and predator abundances were varied and the functional
response was measured as a function of both variables. Arditi & Akçakaya …tted a
model based on the Hassell & Varley (1969) function (1.7),
N
P m N
(1.11)
.
g(N, P ) = g
=
Pm
1 + P m hN
Note that this Hassell-Varley-Holling functional response is prey-dependent for m =
0 while ratio-dependent for m = 1. Actual estimation was done in two steps, …rst
estimating for each predator density the attack rate a = P m by regressing the usual
Holling type II model (nonlinear regression, taking prey depletion into account by use
of the random predator equation, Rogers 1972),
g (N | P ) =
aN
,
1 + ahN
then estimating m by linear regression on log-scale using the relation (1.7). These measurements yielded estimates of m that are mostly closer to 1 than to 0, thus indicating
that the predator isoclines should not only be slanted, but that ratio dependence is
closer to real dynamics than prey dependence.
C. Jost
Identifying predator-prey models (PhD Thesis)
11
Figure 1.1: Predictions of the predator equilibrium abundances in the chained
vessel experiments described in Arditi et al. (1991b), for a ratio-dependent model
(left) and a DeAngelis-Beddington type model (right). N is the algae concentration and P the number of cladoceres. Ison and Isop are prey-isocline and
predator-isocline respectively. D represent cladocere individuals. The …rst vessel
equilibrates as indicated in (a) and (b), letting N1? ‡ow into the second vessel (c)
and (d) and so on. With both models Pi? decreases in a geometrical way.
1.2.3
Mechanistic derivations of ratio dependence
One of the most frequent criticisms of ratio-dependent theory concerns its introduction
as a phenomenological model (Murdoch et al. 1998; Abrams 1994). Originally, Arditi
& Ginzburg (1989) gave several possible mechanisms that can lead to ratio dependence
(different time scales of feeding and population dynamics, predator interference, refuge
for prey, non-random search, general aspects of heterogeneity) but without elaborating
on them. Some of these hypothetical causes have recently been analysed mathematically
and it has been shown that they can indeed lead to ratio dependence. Michalski et al.
(1997) and Poggiale et al. (1998) studied systems where the prey have a refuge patch
and migration of prey between the refuge and the patch with the predator occur at a
slower time scale than the population dynamics in the mixed patch (where predation
is assumed to occur in a prey-dependent way). This model can be aggregated mathematically to a one-patch predator-prey model and results in a linear ratio-dependent
functional response. Cosner et al. (in press) studied different ways of spatial organisation of predators during feeding activity. Depending on the geometry of the predator
distribution (evenly distributed, patchy, aligned), functional responses of several types,
including ratio-dependent, can result.
We can derive ratio dependence also by taking directly the original functional re-
12
Identifying predator-prey models (PhD Thesis)
C. Jost
sponse proposed by Beddington (1975),
g(N, P ) =
aN
,
1 + ahN + cP 0
(1.12)
where c is the product of predator encounter rate and interference time and P 0 = P uP
is one unit predator (uP ) less than the total predator abundance. Beddington derived
this form by letting the predator either search for prey, meet a prey and consume (handle) it or meet a predator and engage in interference (during some constant interference
time). Although there are some problems in his reasoning (Ruxton et al. 1992), this
functional response is often cited in the literature, usually replacing P 0 by P for simplicity. However, keeping P 0 and rearranging (1.12) gives
g(N, P ) =
(1
aN
.
cuP ) + ahN + cP
For c = 0 this becomes the usual Holling type II functional response, while for c = 1/uP
we obtain a ratio-dependent functional response. The Beddington-DeAngelis model is
thus a true intermediate model like the Hassell-Varley-Holling model (1.11).
Additional mechanistic derivations in the microbiological context are mentioned in
chapter 6: Fujimoto (1963) derived ratio dependence based on enzyme kinetics, while
Characklis (1978) based his reasoning on saturation kinetics applied to mass transfer
limited growth.
1.2.4
Alternatives to predator dependence
Before concluding this section, I should discuss possible alternatives to predator dependence in the functional response. Going back to the general predator-prey system
(1.3), we see that slanted predator isoclines, as they are suggested by the empirical
evidence of correlated equilibria along a gradient of enrichment, could also be explained
by density-dependent predator mortality rates (P ) (Gatto 1991; Gleeson 1994). The
idea of such a density dependence has been introduced many times in predator-prey
or more general models (DeAngelis et al. 1975; Steele & Henderson 1981; Steele &
Henderson 1992; Edwards & Brindley 1996), see also Table A.4. While in predator-prey
models this idea indeed gives the desired correlation between equilibrium abundances,
in 3-level food chains with density-dependent mortality at the highest level there can
be negative or positive correlation between the lower two trophic levels (see section
4.B). More complex patterns are to be expected with longer food chains. Predator
dependence in the functional response thus offers a more ‘natural’ explanation since
this feature yields correlated equilibrium abundances of all trophic levels independently
of the speci…c parameter values. Furthermore, the functional response data analysed in
Arditi & Akçakaya (1990) demonstrate directly predator dependence of the functional
response, while I could not …nd such experimental evidence for a density-dependent
mortality rate.
Another alternative is to abandon the general framework (1.3), in particular the
conservation of mass principle. A prominent predator-prey model of this type goes back
C. Jost
Identifying predator-prey models (PhD Thesis)
13
to Leslie (1948). This predator-prey model conserved the functional response in the
prey equation, but used a logistic type form in the predator equation,
N
dN
= r1 (1
) g(N )P
K
dt
dP
P
= (r2 b )N
dt
N
g(N ) = aN
(standard Lotka-Volterra interaction).
(1.13)
An extension is to use a Holling type II functional response (1.6), giving a predatorprey system analysed by May (1975) and by Tanner (1975) and therefore called in
the literature either the Leslie-May model or the Holling-Tanner model. Two features
render this system interesting: First, the existence of limit cycles where the populations
do not come close to extinction, and second, a more varied link between system stability
and enrichment. The idea of density dependence based on the ratio predator by prey
has been generalised to food webs and is quite prominent under the name logistic food
web theory (Berryman et al. 1995). In the literature these approaches are also often
referred to as ratio-dependent models (Freedman & Mathsen 1993; Hsu & Huang 1995),
but referring here to the ratio predator per prey. However, Ginzburg (1998) argues that
conservation of mass is an important feature of food web models and I will discuss in
this thesis only models that comply with this principle.
1.2.5
Ratio dependence: state of the art
The results reviewed in the preceding pages show that predator dependence in the
functional response is a frequent feature of laboratory and natural predator-prey populations, both for theoretical and experimental reasons. Or, the vertical predator isocline
predicted by prey-dependent functional responses in framework (1.4) has experimentally
been proven wrong in many cases. However, many alternatives exist (Hassell & Varley
1969; Hassell & Rogers 1972; Beddington 1975; DeAngelis et al. 1975; Gatto 1991), and
ratio-dependence is only one possibility, although admittedly a very parsimonious one
that can easily be generalised to food chains and food webs while remaining mathematically tractable. This latter argument is in fact a very strong one, since modelling as
a tool in applied ecology (natural resource management, conservation biology, environmental impact assessment, etc.) should be simple to apply with few parameters that
need to be estimated, otherwise the proposed models won’t be used by practitioners.
Two parameters for the functional response represent in this context already an upper
limit, I am not aware of any study in applied ecology using more complex functional
responses. Nevertheless, while the reported experiments (Arditi et al. 1991b; Arditi
& Saïah 1992) and the correlated plankton abundances along a gradient of richness
demonstrate that the predator isocline is not vertical in many cases and that spatial
heterogeneity increases the slope of the predator isocline, both results do not allow to
distinguish between general slanted predator isoclines (Hassell & Varley 1969; DeAngelis et al. 1975) and a predator isocline going through the origin as predicted by the
ratio-dependent model (see also Figure 1.1). Arditi & Akçakaya (1990) and McCarthy
et al. (1995) are the only papers with indications that ratio-dependent models are in
14
Identifying predator-prey models (PhD Thesis)
C. Jost
many systems closer to reality than prey-dependent models (estimates of the interference parameter m being mostly closer to 1 than to 0). However, since this parameter
m appears in a nonlinear context, being closer to 1 than to 0 does not mean necessarily
that the predator isocline in the range of observed abundances is better approximated
by a diagonal rather than by a vertical line (see Figure 1.2 and 4.2).
P
1.2
1
0.8
0.6
0.4
0.2
0.2
0.4
0.6
0.8
1
N
Figure 1.2: The predator iscoline for different values of m in system (1.11). The
grey shaded area indicates the range of values of m found by Arditi & Akçakaya
(1990). For lower values of m the vertical predator isocline might be a better
approximation than the diagonal predator isocline.
Despite all the empirical and theoretical evidence against prey-dependent predatorprey models, ratio dependence is still far from being accepted as a valid theoretical
framework for modelling predator-prey systems or more complicated food chains and
food webs. While some scientists regard it as complete nonsense (Abrams 1994; Abrams
1997) others are afraid that it might de‡ect attention from more general forms of predator dependence (Murdoch et al. 1998). Still others criticise that predators can potentially persist at arbitrary low resource levels (Yodzis 1994). However, this last argument
also applies to logistic growth models, but nobody would argue that logistic growth (1.1)
must be abandoned, e.g., in favour of its counterpart with Allee effect (Allee 1931) (see
section 2.3.1). Furthermore, there is no experimental evidence that disappearance of
predators with decreasing resources is a deterministic rather than a stochastic result.
Demographic stochasticity or mutational meltdown (accumulation of deleterious mutations in small populations due to the weaker power of selection in comparison to genetic
drift, Lynch et al. 1995) will also show this disappearance of predators at low resource
levels. Actually, the reasoning behind Yodzis’ criticism has best been developed by
Oksanen et al. (1981) and leads to the prediction that richer environments permit the
existence of longer food chains. However, Pimm (1991, p. 224) shows that there is no
evidence for such a correlation between primary productivity and food chain length.
In sum, there remains work on several levels to be done. First, prey-dependent
and ratio-dependent models have to be compared to data on an equal footing, e.g.,
C. Jost
Identifying predator-prey models (PhD Thesis)
15
by statistical comparison tools (nonlinear regression in a likelihood framework) or by
developing experiments that can disprove both types of isoclines, vertical ones and those
that go through the origin. Second, there are still some features in ratio-dependent
models that are mathematically poorly understood (dynamic behaviour at the origin
where the models are not de…ned). A thorough mathematical understanding of a model
can indicate both weaknesses and potentially strong features. Third, other possibilities
of including predator dependence in an equally simple way as in ratio dependence have
to be studied and confronted with the ratio-dependent approach. The interference
parameter m proposed by Hassell & Varley (1969) is such a candidate, the functional
response proposed by Ashby (1976) (see Table A.3) another one. In this thesis, I will
contribute mainly to the …rst and to the second task. Other possibilities of predator
dependence are not analysed because they require more model parameters to have the
same range of dynamic behaviours (see section 2.3.1 and chapter 8).
1.2.6
A short digression: to what density does ‘density dependence’ refer in the functional response?
There is some confusion in the literature about the use of the term density dependence
in the context of predator-prey models since density may refer to prey or predator
abundance. A reasonable approach has been used by Ruxton and coworkers who call
the case of predator-dependence a ‘density-dependent’ functional response (Ruxton &
Gurney 1992; Ruxton 1995), because the functional response is intrinsically bound
to the predator (no functional response without predator) and also because density
dependence indicates in its original sense a detrimental character concerning growth of
an organism. However, other authors refer to prey density when using the term ‘densitydependent predation’ (Sih 1984; Trexler 1988; Moore 1988; Mansour & Lipcius 1991;
Possingham et al. 1994). This usage goes actually back to the founder of the terms
functional and numerical response, Solomon (1949, p.16): “To be density-dependent, the
enemy must take a greater proportion of the population as the host density increases”.
Solomon himself refers to Varley (1947) and Nicholson (1933) for this usage. Despite the
historical primacy I consider this to be wrong usage of the term. To avoid any confusion
one might call a functional response that only depends on prey density a prey-dependent
functional response. When this function also depends on predator density one can refer
to a predator-dependent functional response.
16
Identifying predator-prey models (PhD Thesis)
C. Jost
Chapter 2
A case of model selection
To compare prey-dependent models to ratio-dependent models in the full generality included in these two classes of models is a difficult, if not an impossible task. A possible
approach is to take a parametric representative of each class, both of the same complexity (in the sense of having the same number of parameters and of having the same
range of possible dynamic behaviours), and then confront these two predator-prey models qualitatively (stability analysis, reaction to parameter changes, temporal dynamics)
and quantitatively (nonlinear regression, likelihood) to ecological data. A better agreement between one of the models and the data can then be interpreted that the strength
of predator dependence in the functional response is smaller or larger. The data are
either qualitative temporal dynamics of a predator-prey system (plankton dynamics in
freshwater lakes, section 2.2) or time-series data from simple protozoan predator-prey
systems, structurally more complex arthropod systems and complex freshwater plankton
systems (section 2.5).
2.1
The candidate predator-prey models
I will analyse two models. Both have logistic growth of the prey (in the absence of
a predator) and a constant predator mortality rate. The prey-dependent functional
response is of the Holling type II form (1.6), while the ratio-dependent model is also of
this form, but with the ratio N/P (prey per predator) as its argument,
dN
dt
dP
dt
N
)N
K
=
r(1
=
eg (N, P )P
g(N, P )P
P
(2.1)
aN
1 + ahN
g(N, P ) →
N/P
.
1 + hN/P
Both models have six parameters. As will be shown in chapter 4 they have both three
possible equilibria and the dynamics can be either stable coexistence, unstable coexistence (limit cycle), extinction of the predator only or extinction of both populations
17
18
Identifying predator-prey models (PhD Thesis)
C. Jost
(the latter exists in the prey-dependent model only approximatively, i.e., large amplitude
limit-cycle oscillations bringing the dynamics very close to extinction). These dynamics
are also found in the ecological time-series with which I will confront the models.
Density dependence in the prey growth function is required for two reasons, …rst,
that the prey population does not explode in the absence of predators, second, because
malthusian growth combined with a bounded functional response would always yield
an unstable non-trivial equilibrium. There are alternative growth functions (Gompertz
1825; Rosenzweig 1971) and we chose the logistic one simply because of its general
acceptance. Both functional responses are bounded (saturation of the consumption), an
ecologically uncontested feature. There are prey-dependent alternatives to the Holling
type II functional response that have also two parameters (Ivlev 1961; Watt 1959)
and we choose the Holling type II function simply because it is the most widely used
and best understood function. There exist also predator-dependent alternatives to the
ratio-dependent functional response, such as the linear Hassell-Varley type functional
response,
g(N, P ) = P
m
N,
or the form proposed by Ashby (1976) or Hassell & Rogers (1972),
g(N, P ) =
aN
1
.
1 + ahN P
While both have only two parameters like the ratio-dependent model, they both have
the drawback that the functional response is unbounded for P → 0 and that in the
framework (2.1) they cannot have an unstable non-trivial equilibrium or stable limit
cycles. We can therefore safely argue that the Holling type II like ratio-dependent
functional response is the only simple predator-dependent functional response whose
dynamic behaviours are comparable to the ones with a Holling type II (prey-dependent)
function.
2.1.1
Note on discrete versus continuous models
The two candidate models developed in the preceding section are both continuous time
models (differential equations). Such models apply to large populations that have complete overlap between generations. The opposite extreme to this are populations that
are made up of a single generation, with no overlap between generations (e.g., annual
plants). Such populations should evolve in discrete steps,
Nt+1 = F (Nt , Pt )
Pt+1 = G(Nt , Pt )
(see comments to such models in May (1976a) or in Begon et al. (1996a)). The ratiodependent approach has rarely been used in discrete models, the only example to my
knowledge is Carpenter et al. (1993) and Carpenter et al. (1994) (they used discrete
models for technical reasons in their time-series analysis). The real data used in this thesis for model comparison can all be characterised by overlapping generations with large
C. Jost
Identifying predator-prey models (PhD Thesis)
19
populations, they therefore conform better to continuous time models. For this reason
I only used and compared differential equations, which is numerically more demanding
but easier for the theoretical aspects. The usefulness of ratio-dependent approaches in
discrete models remains so far unexplored.
2.2
Qualitative comparison of models
This section will summarise the ideas and results of the analysis developed in chapter
4.
The best known temporal dynamics from non-laboratory systems probably come from
plankton in freshwater lakes of the temperate zone. Studying the detailed dynamics of
phyto- and zooplankton in 24 lakes of the temperate zone Sommer et al. (1986) developed a verbal model (PEG-model, PEG standing for Plankton Ecology Group) of these
dynamics in 24 statements that can be grossly summarised as in Figure 4.1. In early
spring both phytoplankton and zooplankton start growing from the low overwintering
levels. Phytoplankton reaches a very high spring peak, followed by a strong decline to
very low levels due to overgrazing by herbivorous zooplankton and nutrient depletion.
This decline is called the clear water phase (CWP) because the water appears very
clear due to the low phytoplankton levels. After this decline phytoplankton increases
again to the spring peak levels, dominated mostly by inedible species (…lamenteous,
with defence structures, toxic), and remains at these levels during the summer until decreasing temperature and light engender a population decline in autumn. Zooplankton
dynamics follow these dynamics with some lag with less distinct maxima and minima.
The observed dynamics are quite independent of the trophic state of the lake, only
the amplitudes increase and the spring peak is reached earlier with increasing nutrient
status.
2.2.1
Interpretation in the predator-prey context
These dynamics from spring to autumn are interpreted in the PEG-model as the result of
trophic interactions. The simplest possible way is to consider phytoplankton as prey and
zooplankton (mostly Daphnia sp.) as predators and the temporal dynamics correspond
to transient dynamics starting with low initial conditions and reaching the equilibrium
state during summer after some strong oscillations. Chapter 4 explores whether the two
models under consideration in this thesis (2.1) can explain these dynamics and if there
are differences between their predictive power.
In a …rst step the global dynamics of the two predator-prey systems are analysed.
A …rst observation concerns the prey equilibrium and how it changes with enrichment
(eutrophication, modelled by carrying capacity K ): in the prey-dependent model the
prey-equilibrium is independent of carrying capacity, while the ratio-dependent model
predicts a proportional relationship. Observations of natural lakes show a strong correlation between nutrient status and average phytoplankton abundance (McCauley et al.
1988; Mazumder & Lean 1994), thus being in accordance with the ratio-dependent
20
Identifying predator-prey models (PhD Thesis)
C. Jost
model. To study the transient dynamics we …rst construct the isoclines. Both models
have a humped prey isocline and a straight predator isocline, the difference being that
the prey-dependent model has a vertical predator isocline and the ratio-dependent model
a diagonal one (Figure 4.2). There are maximally three equilibria: the origin (0, 0), the
system without predators (K, 0) and a non-trivial equilibrium. This section focuses on
the last (non-trivial) equilibrium which should be locally stable and reached with oscillations to conform to the dynamics of the PEG-model. After non-dimensionalisation
the regions in parameter space that permit such dynamics are identi…ed, see Figure
4.3. The main difference between the models is, again, the reaction to enrichment (increasing K): destabilization in the prey-dependent model, only increasing amplitudes in
the ratio-dependent model. This makes the ratio-dependent model conform better with
the observed dynamics that did not show any sign of destabilization in more eutrophic
lakes. However, the ratio-dependent model can explain the observed earlier attaining
of the spring peak in eutrophic lakes (as described in the PEG-model) only if with
eutrophication the growth rate r also increases.
Despite this slight advantage of the ratio-dependent model, neither model can reproduce the PEG-model dynamics: either there is a strong initial oscillation (causing
the CWP), but then the system continues oscillating during the summer, or the system
stabilises in early summer but then there is no CWP (Figure 4.5). Apparently, the two
predator-prey systems are too simple to explain the dynamics.
McCauley et al. (1988) suggested that several parameters change during the season:
growth rate r increases due to higher water temperatures,
attack rate a () decreases due to an increasing proportion of inedible algae,
predator mortality increases due to higher predation (…sh larvae) in the later
season.
These potential parameter changes suffice in both models to generate the desired dynamics (see Figure 4.6 for an example).
We can conclude that both models are too simple to explain the dynamics of the
PEG-model if parameters are …xed, but that both models can explain them if one or
more parameters are allowed to change during the season. The only distinguishable
feature in the models is the reaction to enrichment (if we translate this enrichment as
only increasing carrying capacity K ). In this case both models contain predictions that
are incompatible with the PEG-model: the prey-dependent model cannot explain the
increasing phytoplankton equilibrium while the ratio-dependent model cannot explain
the earlier spring peak (or generally the changing period in the oscillations). However,
if enrichment also increases r then both models agree with the PEG-model.
2.2.2
Adding a trophic level
The results mentioned in this subsection are not part of the article in chapter 4. They
form a preliminary assessment how the work of chapter 4 could be continued.
C. Jost
Identifying predator-prey models (PhD Thesis)
21
The suggested seasonal parameter changes in the previous paragraph are mostly due
to changes in the in‡uence of another trophic level on the phytoplankton-zooplankton
system (a top predator eating Daphnia, increasing abundance of inedible algae). Instead
of changing the parameter we can also directly model this trophic level. Adding a top
predator C gives the system
dN
dt
dP
dt
dC
dt
= f (N )N
g1 (N, P )P
= e1 g1 (N, P )P
g2 (P, C )C
= e2 g2 (P, C )C
C.
(2.2)
Bernard & Gouzé (1995) developed a method to study the possible succession of extrema
for this type of food chain models. Following the dynamics of the PEG-model and
marking each trophic level with a + if it increases and with a
if it decreases then
we can depict the PEG-model dynamics in the following succession of -vectors (see
Appendix D for details),
’
+
+
- +
- +
- +
- +
$
sign(N˙ )
- + := sign(P˙ )
sign(C˙ )
&
%
Using the method by Bernard & Gouzé (1995) we can analyse prey- and ratio-dependent
food chains if they can show this succession of events. This analysis shows that both
food chains have exactly the same succession of extrema and can predict the above
depicted succession (see Appendix D for details).
We can make a more quantitative analysis by directly parameterizing system (2.2)
with logistic growth of the prey and prey-dependent or ratio-dependent functional responses (of the form given in 2.1) and demonstrate that there exist parameter values that
produce the observed dynamics. This represents some kind of model validation since
we check whether the proposed model can show the observed dynamics. We assume the
top predator in this chain to increase in abundance during the season. Figure 2.1 shows
such an example for both types of the functional response. In this 3-level food chain
the ratio-dependent model retains its advantages against the prey-dependent model,
notably the correlated prey and predator equilibria along a gradient of richness.
We can also explicitly model the two types of prey, edible (Ne ) and inedible (Ni ),
that are in competition one with each other,
Ne
dNe
= re (1
)Ne ce Ne Ni
Ke
dt
dNi
Ni
= ri (1
)Ni ci Ni Ne
dt
Ki
dP
= eg (Ne , P )P P,
dt
g(Ne , P )P
(2.3)
22
Identifying predator-prey models (PhD Thesis)
C. Jost
Figure 2.1: Numerical example of a 3-level food chain ‘prey-predator-top predator’
with prey-dependent (a) or ratio-dependent (b) functional response. The characteristic elements of the PEG-model are well preserved. ————: prey dynamics;
– – –: predator dynamics; — — —: top predator dynamics.
the functional response being again of the forms (2.1). A reasonable scenario for this
system is that Ne is a superior competitor to Ni (Agrawal 1998) but does not exclude it
due to losses by predation. Figure 2.2 shows a numerical realisation of such a system.
Again, the dynamics of the PEG-model are clearly reproduced. With this scenario
both types of functional response predict a correlation between total prey and predator
equilibria along a gradient of richness.
(a)
time
(b)
time
Figure 2.2: Numerical example of the dynamics of a system with edible prey, inedible prey and a predator. Both the prey-dependent (a) and the ratio-dependent (b)
model show the characteristics of the PEG-model. ————: edible prey dynamics;
– – –: inedible prey dynamics; — — —: predator dynamics.
2.2.3
Conclusions
In summary, neither the prey-dependent nor the ratio-dependent predator-prey models
composed of phytoplankton and zooplankton are able to predict seasonal dynamics of
these two levels except if parameters are allowed to change during the season (in which
case both models can predict the dynamics). Nevertheless, the ratio-dependent model
seems to be more realistic in view of the following properties: 1. both phytoplankton
and zooplankton biomasses at equilibrium increase with increasing productivity and
2. there is no effect of productivity on the stability of the system. To obtain the
expected pattern, with a clear water phase and a stable equilibrium rapidly reached
in the summer, the following modi…cations of the two-level ratio-dependent model are
C. Jost
Identifying predator-prey models (PhD Thesis)
23
proposed: 1. parameters of phytoplankton and zooplankton change during the season,
2. introduction of a third level (carnivorous zooplankton or …sh), 3. introduction of
phytoplankton mortality other than zooplankton grazing (sedimentation, lysis).
2.3
Boundary dynamics: do they matter?
This section summarises the ideas and results of the analysis developed in chapter 5.
In the previous section we have compared our two predator-prey systems (2.1) to data
reaching a stable non-trivial equilibrium. The ratio-dependent model also offers deterministic extinction as a possible outcome (Arditi & Berryman 1991). However, a
ratio-dependent model is not directly de…ned at the origin since there occurs a division
by 0. Nevertheless, if we assume a bounded functional response (as in systems 2.1)
then the origin is indeed an equilibrium point, we only cannot apply standard stability
analysis to this point, making that the dynamics around the origin have been badly
understood for a long time. These badly understood dynamics have been repeatedly
used to criticize the ratio-dependent concept (Yodzis 1994; Abrams 1997) and wrong or
unprecise mathematical results have been published (Getz 1984; Freedman & Mathsen
1993).
In collaboration with Ovide Arino I have analysed the analytic behaviour around
the origin in detail. The results are best explained graphically. Figure 5.1 shows possible isocline portraits of the ratio-dependent model (2.1). Cases 5.1a and 5.1c offer
no mathematical problems, it is the situation in 5.1b that is badly understood. By a
transformation of state variables (studying systems (N/P, P ) and (N, P/N ) instead of
system (N, P )) one can show that for certain parameter values the origin indeed behaves
like a saddlepoint (Figure 5.2). However, with increasing predator efficiency (visualised
by an increasing slope of the predator isocline) the origin becomes attractive for trajectories where ‘N goes faster to 0 than P ’. This already happens at parameter values
where the non-trivial equilibrium is still locally stable (Figure 5.3), thus dividing the
phase space into two basins of attraction. The existence of such trajectories is proven
for general ratio-dependent models, only using that the functional response is increasing
and bounded and the per capita prey growth function is a decreasing function of prey
abundance. Further increasing predator efficiency can lead to a destabilization of the
non-trivial equilibrium, giving rise to a stable limit cycle (Figure 5.4). However, the
existence of this limit cycle is still not analytically proven. The problem is that the origin becomes attractive before the non-trivial equilibrium becomes unstable. Therefore
the construction of an invariant set, as required for the application of the PoincaréBendixson theorem to prove the existence of a stable limit cycle, would require precise
knowledge of the separatrix between the two basins of attraction. Further increasing
predator efficiency leads to a homoclinic bifurcation (detected numerically) after which
the origin becomes attractive for all initial conditions except the (unstable) non-trivial
equilibrium itself (Figure 5.5).
Gragnani (1997) raises the point that this deterministic extinction is not a generic
behaviour, arguing that if we take an intermediate model of the DeAngelis-Beddington
24
Identifying predator-prey models (PhD Thesis)
C. Jost
type (1.12) then the origin remains a saddle point for any parameter value but the
one that corresponds to the ratio-dependent model. The author does not tell that
this argumentation rests completely on the (arbitrary) choice of an intermediate model
(by the way, the author also completely missed the homoclinic bifurcation, i.e., the
global attractivity of the origin while there exists a non-trivial equilibrium). If we
take the Hassell-Varley-Holling functional response (1.11) as an intermediate model the
picture changes somewhat. Callois (1997) studied this model in detail and showed that
deterministic extinction occurs for m 1, while the origin is a saddle point for m < 1.
It therefore indeed seems that deterministic extinction is not generic for m 1.
2.3.1
Some other models that offer deterministic extinction
The following two sections are not detailed in chapter 5. They contain additional ideas
about deterministic extinction and biological control.
Extinction of both prey and predator is usually explained a the result of stochastic
events. However, the consistency of extinction occurring in laboratory predator-prey
systems (Gause 1934; Luckinbill 1973; Veilleux 1979) suggests that it can also happen
as a deterministic outcome. Besides the ratio-dependent model analysed above I only
found two other (simple) predator-prey systems that offer this type of extinction. The
…rst one is a model with a strong Allee effect (population rate of change becoming
negative below a certain population threshold value , Allee 1931) in the prey dynamics
and Lotka-Volterra functional response, e.g. (from Hastings 1997)
dN
dt
dP
dt
= r(1
= eaN P
N
N
)N (
K
1)
K
K
aNP
P.
Note that a general Allee effect only means that the per capita population growth rate
is maximal not at the origin but at some positive population density (Edelstein-Keshet
1988). The strong Allee effect I refer to means that this per capita growth rate must
even become negative a low densities (see Figure 2.3). Only this strong Allee effect
permits extinction of the prey. The second model takes logistic prey growth and a
predator-dependent functional response that has been proposed by Hassell & Rogers
(1972) or Ashby (1976),
dN
dt
dN
dt
N
aN
1
P
)N
K
1 + ath N P
aN
1
P P.
= e
1 + ath N P
= r(1
This model is completely donor controlled and can have two non-trivial equilibria (one
of them stable), in which case the origin becomes attractive for all initial conditions
with sufficiently low prey abundance.
C. Jost
Identifying predator-prey models (PhD Thesis)
25
Figure 2.3: Types of prey per capita growth functions f (N ) in the model dN/dt =
f (N )N : (a) the logistic growth, (b) a weak Allee effect, K is the only stable
equilibrium and (c) a strong Allee effect, K and 0 are stable equilibria. K is the
carrying capacity, N the prey density.
2.3.2
Is there a biological control paradox?
The analysis in chapter 5 revealed another interesting feature of ratio-dependent models: stable coexistence of prey and predator is possible with prey equilibrium values
arbitrary close to 0. This is in contrast to prey-dependent theory where such low prey
equilibrium values either lead to a destabilization of the system or require very high
predator abundances (Arditi & Berryman 1991). In reply to Arditi & Berryman (1991),
Åström (1997) pointed out that a nonlinear density dependence in the prey growth
function in the form of -logistic growth (in ecology introduced by Gilpin & Ayala 1973,
but …rst used by Richards 1959),
!
N
f (N ) = r 1
,
K
could stabilise predator-prey systems with low prey-equilibria. However, few estimates
of the parameter exist. Åström (1997) cites some tentative studies that indicate this
value to be below 1 for insect types of prey and above 1 for mammalian types of prey.
Only < 1 has the stabilising effect on the prey equilibrium described by Åström.
However, a small reduces also the population growth rate at low population levels,
which is somewhat in contrast to features of a typical pest.
Actually, the biological control problem has been addressed earlier by Beddington
et al. (1978). They de…ned so-called q-values, that are the ratio of the (stable) prey
equilibrium coexisting with the predator divided by the prey equilibrium in absence of
the predator. Small q-values (<0.01) thus indicate a successful biological control of the
prey by its predator. Beddington et al. (1978) reported q-values ranging in the …eld from
0.002 to 0.03 and in the laboratory from 0.1 to 0.5. They explored several possibilities
how predator-prey or host-parasitoid models could be adapted to accommodate these
observations:
sigmoidal functional responses had no in‡uence on q,
inclusion of mutual interference in a DeAngelis-Beddington way (1.8) gave q-values
down to 0.1,
26
Identifying predator-prey models (PhD Thesis)
C. Jost
incorporating spatial aggregation of parasitoids with heterogeneous host distributions made q-values below 0.1 possible.
In an earlier paper Free et al. (1977) had shown that spatial aggregation, observed
from a global perspective, leads to reduced attack rates a (like predator interference,
see equation (1.7)), a phenomenon that they termed ‘pseudo-interference’. Beddington
et al. (1978) therefore modelled the last item in the list above with Hassell-Varley’s (1.7)
interference model, based on estimates for the interference parameter m ranging from
0.1 to 0.8, and they obtained q-values down to 0.004. Using the Hassell-Varley-Holling
(1.11) functional response Arditi & Akçakaya (1990) had shown that the parameter m
could even be higher, thus allowing for very low q-values. As already pointed out, the
Hassell-Varley-Holling model becomes ratio-dependent for m = 1, thus proposing this
model as a very rough description of biological control problems.
Let us identify the region in parameter space of a ratio-dependent model that shows
such low stable prey equilibrium levels. To simplify this analysis we use the dimensionless form of the ratio-dependent model that was introduced in chapter 4,
dN
dt
dP
dt
= R(1
=
N
)N
S
SN
P
P + SN
SN
P
P + SN
QP,
with R = rh/e, S = h/e and Q = h/e. S is the equilibrium value of the prey in the
absence of a predator. We therefore want to identify all parameter combinations that
give a low and stable prey equilibrium N ? ,
N ? < S.
Some algebra shows that this is possible for all S that ful…l
Q Q2 + R
R
R
,
< S < min
(1 )
1 Q
1 Q
1 Q2
Figure 2.4 illustrates the region in parameter space that ful…ls this condition. Approximatively, we therefore need that S 1 RQ , which means in the original parameters
er
.
e h
This condition can be veri…ed with real data. Bernstein (1985) studied the predator-prey
system Tetranychus urticae (prey) with Phytoseiulus persimilis (predator) and parameterized a predator-prey system with each population structured in three life stages.
We can use his estimates to parameterize the ratio-dependent model (in parentheses
is the name of the parameter used in the original publication): r = 5.34d 1 (r1 ), =
0.0433d 1 (2 ), e = 0.159 (min {rm , r2 }) (eggs laid per egg eaten) and h 0.1d 1 (Ki /j
or bi /i ). These values give
5.49
cm2
m2
= 0.00055 .
d
d
C. Jost
Identifying predator-prey models (PhD Thesis)
27
(1- )R/(1-Q)
Figure 2.4: The hatched region in the R S parameter space (for a …xed Q < 1)
indicates the parameter values for which the predator reduces the prey to a stable
equilibrium a fraction below the levels the prey would attain in the absence of
predators. The vertical hatching indicates the area where the origin behaves like
a saddle point, while in the area with horizontal hatching the origin is attractive
for certain initial conditions.
This value is close to the = 0.0002m2 /d reported in (Hassell 1978b) for this predatorprey system. Of course, this is only a preliminary result, but it is encouraging that the
ratio-dependent model may be useful in biological control.
It might also be that predator-prey systems in biological control do not show a
stable equilibrium but some metapopulation dynamics with frequent local extinctions.
Luck (1990) reported (based on data from Murdoch et al. (1985)) that “of 9 successful
biological control projects, only 1 showed to have a stable N ? , the others could be
explained by local extinction”. Again, such local extinctions are possible with a ratiodependent model describing local dynamics. The condition on parameters is similar to
the one found above,
S
R
1
Q
or S >
R
1
Q
,
and the types of extinction are illustrated in Figures 5.5 and 5.1c.
2.4
Microbiologists did it
This section summarises the ideas and results of the analysis developed in chapter 6.
In fact, while Arditi & Ginzburg (1989) were the …rst to propose ratio dependence as a
general concept (with all its implications to bottom-up and top-down concepts), Getz
28
Identifying predator-prey models (PhD Thesis)
C. Jost
(1984) published a speci…c ratio-dependent model several years before them. The very
idea that consumption (and with it growth of an organism) is a function of the ratio
consumer by resource can be found in the microbiological literature already in Cutler &
Crump (1924). Most interestingly, a speci…c ratio-dependent function of the form (1.10)
has been proposed in microbiology as a growth function (same concept as the functional
response) by Contois (1959). This function has since then served as an alternative to the
microbiological equivalent of the Holling type II function, Monod’s (1942) well known
growth function. With the help of Science Citation Index I checked all papers citing
Contois’ work in the last 40 years to detect results that might be useful in the ecological
context.
Most experimental work has been done in the framework of chemostats that are
usually described by
dN
dt
dP
dt
= D(N0
N)
= (N, P )P
Y (N, P )P
DP
(2.4)
with substrate N , in‡owing substrate concentration N0 , consumer P , dilution rate D
and yield Y . Comparing this equation with our standard predator-prey system (1.4)
shows that there is a subtle difference between the two ways consumption is modelled:
microbiologists start with the growth function (called numerical response in population
ecology) and consider the substrate uptake function (our functional response) to be
proportional to it with yield constant Y , while ecologists start with the functional
response and consider the numerical response to be proportional to it with ecological
efficiency e. There is therefore a scaling difference between the two concepts,
g(N, P ) = Y (N, P ).
However, this difference is of no importance to the results discussed below.
Growth rate (N, P ) is usually modelled with the equation proposed by Monod
(1942),
(N, P ) = max
N
Ks + N
with maximal growth rate max and half saturation constant Ks (note that with a = /K
and h = 1/ we obtain the usual Holling type II functional response (1.6)). With
this growth function, system (2.4) predicts that out‡owing substrate concentration N ?
should only depend on D, max and Ks , but not on in‡owing substrate concentration N0 .
However, it was soon observed that in many experiments out‡owing substrate concentration was proportional to in‡owing substrate concentration. Contois (1959) tried to
accommodate this observation by hypothesising (and providing experimental support)
that the half saturation constant is proportional to in‡owing substrate concentration.
Combining this observation with mass balance principles (see chapter 6 for details) he
derived the form
(N, P ) =
um N
BP + N
C. Jost
Identifying predator-prey models (PhD Thesis)
29
(with constants um and B) which is a particular ratio-dependent form.
The papers that tested this function against Monod’s or other functions (see Table
6.1 for examples) can be classi…ed into two categories,
chemostat experiments with varying in‡owing substrate concentrations and dilution rates, measuring out‡owing substrate concentration,
direct measurements of growth with varying substrate and organism concentrations, then comparing these measurements quantitatively and qualitatively with
different growth functions.
The …rst category showed that Monod’s prediction is correct as long as one works with
a unique strain of organisms growing on pure substrate. Any deviation from these
conditions (mixed substrate or multiple strains/species of organisms) lead to out‡owing
substrate concentrations that are proportional to in‡owing substrate concentrations.
This result can be interpreted that the consumer isocline in chemostats (Figure 6.2)
should be slanted. The second category showed that the growth function is a decreasing
function of consumer density,
d(N, P )
< 0,
dP
under the same conditions mentioned for the …rst category (mixed substrate or multiple strains/species of organisms). However, these results are more precise in that
they identify predator dependence in the growth function to be the likely cause of the
slanted predator isocline, and not some density-dependent consumer mortality as proposed by other researchers. While all these results suggest that we should expect a
slanted predator isocline in natural systems, this slanted isocline is also predicted by
any consumer-dependent growth function, not only by Contois’ equation. As in ecology (section 1.2.5), the comparison between Monod and Contois is not done on equal
footing. Fortunately, there were also direct quantitative comparisons between data and
models, based on statistical criteria such as goodness-of-…t. These showed that in many
cases (wastewater treatment, fermentation processes) Contois’ function …tted better
than Monod’s function, but often the identi…cation was less clear or other functional
forms …tted equally well as Contois’ equation.
We can learn several things from the work done in microbiology:
Prey-dependent (or substrate-dependent) functions such as the one proposed by
Monod seem to be a good description of consumption processes of single species
consumers on a pure substrate.
Whenever the system becomes more complex (several species involved, heterogeneities in space or time) predator-dependent functions (of any form, not necessarily ratio-dependent) are more adequate.
If model selection by …tting models to data is already ambiguous with microbiological data, then we cannot expect more conclusive results when working with
(noisier) …eld population dynamics, or we have to develop more sensitive techniques such as proposed in chapters 7 and 8.
30
2.5
Identifying predator-prey models (PhD Thesis)
C. Jost
Quantitative comparison of models
This section summarises the ideas and results of the analysis developed in chapters 7
and 8.
In the beginning I have opined that existing comparisons between prey-dependent or
ratio-dependent predator-prey models and real data have not been done yet on an equal
footing. The analysis in chapter 4 was already closer to this objective by comparing
predictions of each model with the dynamics observed in freshwater lakes. Here, I will
compare the models (2.1) quantitatively based on their temporal dynamics, by …tting
them directly to time-series data of predator-prey systems. The concept of likelihood
(Edwards 1992) provides the statistical framework to ensure that the comparison is done
on an equal footing. The idea is to identify the likelihood of the model given the data
with the probability of obtaining the data given the model (with model parameters that
maximise this probability, maximum likelihood). This identi…cation is intuitive (see
Press et al. (1992) for some comments) but very useful for model selection. Estimation
by maximum likelihood (ML) has a number of interesting statistical properties: the
estimation is consistent and ML is the estimator with the smallest possible variance
(Huet et al. 1992). However, the distribution of the error has to be known exactly for
these properties to hold. ML estimation is often not very robust against uncertainty
concerning this distribution.
If the models that are compared have the same number of parameters (as is the case
in our candidate models) then the likelihoods also serve as selection criterion. Therefore,
regression of the model to data (with parameter estimates as a byproduct) and model
selection are based on the same criterion. Chapter 7 will show that for applications of
model selection this is an important point.
The concept of likelihood can also be extended to compare models with a different
number of parameters, using so-called information criteria. The best known is Akaike’s
Information Criterion AIC (Akaike 1973; Bozdogan 1987), but others exist, such as
Bayesian information criterion BIC, Mallow’s Cp (see Hilborn & Mangel (1997) for a
discussion and references to all three criteria) or consistent Akaike information criterion
CAIC (Jones 1993). However, since I only worked with models of equal complexity
I mention these criteria only for completeness, they are not pertinent for the present
analysis.
In contrast to the equilibrium experiments in Arditi & Saïah (1992) we approach
the problem of model selection from a dynamic point of view. We investigate whether
time-series data of predator-prey systems can reveal if there is predator dependence
in the functional response or not. The comparison will be done by …tting our equally
complex prey-dependent and ratio-dependent predator-prey models (2.1) to time-series
data, and apply goodness-of-…t (likelihood) as a criterion.
The time-series I use in the analyses come either from the literature (protozoan,
arthropod and plankton predator-prey systems) or are original phyto- and zooplankton
data from lake Geneva. They are characterised, as most ecological data, by small size
(10 - 50 data triplets {time, prey, predator}) and by rather large errors with coefficients
of variation up to 50%. By such reasons one should be extremely cautious in making
C. Jost
Identifying predator-prey models (PhD Thesis)
31
any deduction regarding the biological ‘correctness’ of the used formalism to describe
the biological processes even if some model …ts such data very well. The problem has
been illustrated convincingly for the case of logistic population growth (introduced by
Verhulst (1838)) already very early in the history of mathematical population biology,
namely in 1939 by W. Feller. He chose two other mathematical models that are Sshaped and that have the same number of parameters and showed that both forms …t
equally well (or even better) to real data that where considered at the time to be some
of the best proof that the logistic model has the character of a physical law. Feller
continued to show that these models also predict the same response to a constant per
capita harvesting rate as the logistic model.
More recently, Simons & Lam (1980) demonstrated that the same caution applies to
more complex models such as models describing the dynamics of phosphorous in large
lakes. They pointed out that several choices of the parameter values (with or without
imposed seasonality) gave equally good …ts to one season of data, and thus that the
predictive power of such a (…tted) model is very limited.
These two examples amply show that obtaining a good …t of a model to data is far
from proving the biological correctness of the chosen mathematical process description.
Or, in the words of Cale et al. (1989), ‘multiple process con…gurations can produce
the same pattern’. However, there is a solution, following May’s (1989) advice by
‘generating pseudo-data for imaginary worlds whose rules are known, and then testing
conventional methods for their efficiency in revealing these known rules’. In other words,
I will generate arti…cial predator-prey data, apply the model selection criterion to them
and then check if the correct model has been identi…ed. Only after detection of the
limits in model identi…cation in this simulation analysis will I endeavour to analyse real
ecological time-series data. This analysis is summarised in section 2.5.4.
This same approach has been used in Carpenter et al. (1994) for plankton data, but
the authors used a discrete model with the time steps being the (arbitrary) intervals
between measurements. They also …tted time-series over several years, thus assuming
that parameters such as carrying capacity or mortality rates are the same for every
year. I will reanalyze their data by …tting our models to data from one season only, in
particular to data from the period of spring to autumn where population interactions
may be the driving forces behind the dynamics (Sommer et al. 1986), and not physical
constraints such as temperature or light. I will also …t the models to protozoan timeseries found in the literature, because they show much less variability than whole lake
plankton data and there are many similarities between protozoa and plankton (‡uid
medium, overlapping generations, large numbers etc.). There are also more and more
ecological models that include the microbial loop (eg. Fasham et al. 1990, Azam
et al. 1983, Stone & Berman 1993) based on a large body of empirical evidence for the
important ecological role of the microbial fauna (Berman 1990; Fenchel 1988; Sherr &
Sherr 1991). It will be interesting to see how typical ecological models …t the dynamics
of these organisms.
Finally, this work will also be different from the approach presented by Harrison
(1995). The author used a classical predator-prey time-series from the literature (Luckinbill 1973) and …tted it to 11 different predator-prey models, trying to determine the
key processes necessary to get a qualitatively and quantitatively good …t. Two problems
32
Identifying predator-prey models (PhD Thesis)
C. Jost
arise in his work: (1) he implicitly assumed that a better …t is due to a better description of the individual processes and (2) without any information about the size of the
observation error no statistical test could be used in the comparison that accounts for
the number of parameters in the model, e.g. likelihood ratio tests (Hilborn & Mangel
1997) or Akaike’s information criterion (AIC, Akaike 1973). I will show that already
with modest errors in the data assumption (1) may be wrong and detailed simulation
studies should be done preliminary to such work. Unsurprisingly, the model with the
largest number of parameters gave the best …t in Harrison’s (1995) analysis. While this
kind of study may yield information about details in processes of speci…c systems (e.g.
importance of lags between consumption of Paramecium aurelia and reproduction of
Didinium nasutum in this case) it does not tell whether addition of these details gives
a signi…cantly better …t to the data. I will avoid this kind of fallacy by working only
with models that have the same number of parameters.
2.5.1
A remark on model selection criteria
Model selection in this thesis is always based on likelihood (goodness-of-…t), at least
when working with real data. Another approach was chosen by Bilardello et al. (1993),
who used the joint linearised con…dence region as a comparison criterion. This region is
minimal if the determinant of J ()T J ( ) is maximal (Hosten 1974), where J () is the
sensitivity matrix of the state variables with respect to parameter vector (J ()T J () is
also known as Fishers information matrix, Huet et al. 1992). I did not test this criterion,
but it surely would help detecting problems due to overparameterization (which makes
this matrix close to singular). I encountered such problems with the highest noise
levels in the ‘in silico’ analysis (2.5.4). Sometimes there was very slow convergence
even close to the optimum, at other times several different parameter sets gave equally
good …ts. Both phenomena indicate that too many parameters were …tted to the data.
Another popular criterion for model selection is the standard deviation of the residuals
(Carpenter et al. 1994).
All these criteria can pose the problem that model selection is based on a quantity
that is not minimised while …tting the model to data. In chapter 7 I analysed some
such criteria that were proposed by Carpenter et al. (1994). I often observed that a
model was selected that gave visually a worse …t to the data than the rejected model (see
section 2.5.4). Using only the likelihood concept for both regression and model selection
avoids this problem and is statistically well founded. For this reason I remained in this
concept for the analysis of real data.
2.5.2
The errors that govern our modelling world
I have mentioned in the previous section that my model comparison will be based on a
likelihood approach. To formulate this likelihood I …rst have to formulate a stochastic
model, i.e., a model that explains how random effects come into the data to which I
want to …t my candidate models. These models themselves are deterministic, formulated
with ordinary differential equations (ODE) that, for given parameters, give completely
C. Jost
Identifying predator-prey models (PhD Thesis)
33
predictable results. They form the core of the statistical model to which stochasticities
are added.
A …rst type of stochasticity is pure observation or measurement error: there is some
unknown real population abundance and any estimation of it is a realisation of a random
number according to some distribution function. If we have data with only this type
of error (the dynamics of the population are deterministic) then we …t our model by
…tting the whole trajectory (solution of the ODE with given initial conditions that are
treated like parameters to be estimated) to the time-series data (note that we assume
the time measurements to be exact, without any error). This type of …tting is often
called trajectory …tting. See Figure 8.1a for an illustration of this type of …tting.
A second type of stochasticity is process error: the dynamics of a population over
time are not completely governed by its deterministic component but also by random
in‡uences either from the environment or from inside the population. In the context of
an ODE-model process error leads to a stochastic ordinary differential equation (SODE).
If our system has this type of error, then the deterministic model trajectory diverges
more and more from the actual population abundance, the longer the prediction horizon,
the worse the prediction. The population abundance after some …xed time interval
with some initial abundance (which is known if we measure the abundance without
observation error) is thus a realisation of a random process. To relate in this case
the problem of parameter estimation (and of estimating goodness-of-…t) to a regression
problem in the classical sense, we have to make the assumption that the distribution of
the population abundance after the …xed time interval, given the initial conditions, is
known. Usually, the prediction horizon is one time step ahead (de…ned by the spacing
in the available time-series), but it is also possible to predict s > 1 time steps ahead.
This type of …tting is therefore often called s-step-ahead …tting. See Figure 8.1b for an
illustration of this type of …tting.
Solow (1995) distinguishes a third type of error, so-called parameter error. This
is a special kind of process error due to the stochastic nature of one (or several) of
the model parameters. Solow shows that this type of error can lead to non-stationary
error, depending on the probability density function and the algebraic embedding of
this parameter. In my …tting I did not see a way to distinguish between parameter
error and process error as modelled by a SODE. Furthermore, the algebraic structure
of equations (2.1) makes it difficult to deduce how parameter error translates itself into
process error, an information necessary to adapt the regression process. I therefore only
considered process error in form of a SODE, according to the standard treatment of
errors in time-series (Hilborn & Mangel 1997; Pascual & Kareiva 1996; Dennis et al.
1995).
Statistical theory has mainly developed methods to …t one single type of error, the
problem of considering process and observation error at the same time being much
more difficult (discussion in section 3.2.4). If the researcher can design his experimental
setup so that one of the errors is eliminated or at least reduced to negligible values
(e.g., providing constant environmental conditions in the laboratory and working with
population sizes where demographic stochasticity is unimportant, or, developing a sampling method that has no observation error), then the formulation of the stochastic
model poses no problem. However, this is difficult to achieve, especially when working
34
Identifying predator-prey models (PhD Thesis)
C. Jost
in the …eld. Most time-series data available in the literature or by a researchers own
experiment contain both types of errors. Pascual & Kareiva (1996) thus come to the
conclusion that the researcher has to decide (as a best guess) which error to model and
which error to neglect. Due to its pertinence I will often use their terminology, calling
the case of trajectory …tting ‘observation error …t’ and the case of s-step-ahead …tting
‘process error …t’.
The particular choice of one error type or the other can profoundly in‡uence the
result of a model selection process. However, the sensitivity of model selection to the
particular error type can be tested in a simulation analysis, as proposed by May (1989).
This will be the subject of the next section. We discuss and analyse there also two error
functions proposed by Clutton-Brock (1967) and applied by Carpenter et al. (1994) that
seem to take both types of error into account without increasing the numerical costs
too much. However, these error functions do not …t into the likelihood concept, posing
the problem of what criterion to use for model selection. The next section will illustrate
some consequences of this problem. Another statistical method that takes both errors
into account is the errors-in-variables approach that will be discussed brie‡y in section
3.2.4. I have not used this method because the computing cost is much higher, the
method itself is still in its statistical infancy or at best adolescence and the method
requires independent knowledge of either process error or observation error. However,
the method …ts completely into the likelihood concept and looks promising for future
research.
Note that I assume for observation error …t and for process error …t stationarity and
normality of the error on log scale (constant coefficient of variation CV ). With this
assumption maximum likelihood becomes equivalent to ordinary least squares (Press
et al. 1992), and when I will use the terms goodness-of-…t or likelihood the reader best
thinks of them in terms of least squares. However, the size of this error (CV ) is for
the real ecological times series only approximatively known. Furthermore, as already
mentioned, these time-series contain both types of stochasticity and neglecting one of
them is only a statistical necessity. This means that the calculated likelihoods will
not have an absolute meaning and the information criteria mentioned earlier cannot be
used. For this reason, the comparison can only be performed for models of the same
complexity.
2.5.3
Implementation of …tting algorithms
Before discussing the actual …tting results I should say some words on the implementac
tion of the …tting algorithms. Most algorithms were …rst developed in Mathematica
to get an idea if they work satisfyingly. However, numerical calculations in Mathematica are too slow for the analysis of large numbers of time-series. There does not exist
much customisable software that allows …tting differential equations to data. Fitting
and visualisation are usually separate steps in most software, which further slows down
the …tting process. Therefore I implemented all algorithms in C++, using the MacIntosh Toolbox functions to visualise the …tted model in real time. This immediate visual
control proved to be an indispensable tool to analyse large numbers of data sets. The
numerical algorithms are based on the codes provided with ‘Numerical Recipes in C’
C. Jost
Identifying predator-prey models (PhD Thesis)
35
(Press et al. 1992). The software is obtainable on request from the author (it requires
a Power Macintosh).
2.5.4
An ‘in silico’ approach to model selection
The objective of the analysis detailed in chapter 7 is to explore numerically how the
presence of both observation and process error in time-series data in‡uences model
selection based on goodness-of-…t. For this I created arti…cial time-series that contain
both errors. The deterministic core function is inspired by the dynamics described in the
PEG-Model (section 2.2), i.e., dynamics that reach a stable equilibrium after one or two
large amplitude oscillations. In a …rst step I computed parameters and initial conditions
for both the prey-dependent and the ratio-dependent candidate model (2.1) in a random
procedure. The criteria to accept such a random parameter set are on the one hand
the dynamics described above and on the other hand practical requirements such as
prey and predator equilibria not differing by more than a factor of 100 and predator
dynamics showing at least a …ve-fold variation. For each model 20 such parameter
sets were created. In analogy to replicated experiments in ecology I then simulated 5
replicate time-series with each parameter set, adding process error in the framework
of stochastic ordinary differential equations, ‘sampling’ the resulting time-series at 20
equal intervals (also inspired from the PEG-model, 20 is the typical number of plankton
samples over one season) and adding an observation error to these. The process error is
generated in a standard stochastic process framework (see details in section 7.3.1) with
a stochastic component that is proportional to the current population abundance, thus
simulating a multiplicative or lognormal error type. Observation error is of a lognormal
type with constant coefficient of variation. Time-series with a low observation and
process error (comparable to laboratory protozoan systems) and high observation and
process error (comparable to aquatic mesocosms, maybe even to whole lakes) were thus
created for each model.
Then each candidate model was …tted to each time-series with four methods (or
error functions). Method A assumes that there is only observation error of a lognormal
type, observation error …tting or trajectory …tting. In this setup maximum likelihood
corresponds to least-squares …tting on the log-scale. Method B assumes that there is
only process error of a lognormal type, process error …tting or s-step-ahead …tting. In
this setup maximum likelihood corresponds to conditional least squares on the log-scale.
The prediction horizon s is chosen such that the autocorrelation drops below 0.5 (Ellner
& Turchin 1995). The reason for this is to …t over a time interval where nonlinearities
may be equally (or more important) as linear dependencies. Actually, this criterion
for the prediction horizon is adopted from Ellner & Turchin (1995) and justi…ed there
mainly as an empirical result from experience. Methods C and D are both inspired
from Clutton-Brock (1967) as presented by Carpenter et al. (1994). C corresponds to
weighted conditional least squares, where the weights are computed from the sensitivity
of the prediction to the initial condition, which is the observation s steps previously and
that contains an observation error (see section 7.3.2 for details). The observation error
is assumed to be known independently from process error. D is based on C but is more
similar to a negative log likelihood function in that the scaling factor of the normal
36
Identifying predator-prey models (PhD Thesis)
C. Jost
distribution has been retained (see Carpenter et al. 1994) such that larger weights not
only reduce the importance of the residuals but also add a penalising factor to the error
function. Note that these so-called loss functions C and D (Carpenter et al. 1994)
do not …t into the likelihood concept. Model selection must therefore be based on a
different criterion. I chose the sum of squared residuals (computed on log-scale) as this
selection criterion.
Time-series with low observation and process error were reliably identi…ed (with less
than 5% wrong identi…cations) by methods A, B and D. Method C had close to 15%
wrong identi…cations and was therefore rejected as a useful method for model selection
(see Figure 7.3).
Time-series with high error levels were best identi…ed by A (observation error …t)
with less than 5% wrong identi…cations. B and D had both 10% wrong identi…cations,
but B was better in the sense that if we require a minimal difference of 5% between the
…ts then this method identi…ed more than 95% of the time-series correctly (Figure 7.3).
No such threshold could be found for D.
The problems with C and D are two-fold. First, the algorithms often rather maximised dependence on initial conditions (larger weights) than minimised the residuals.
Second, the model selection criterion often increased while minimising the loss function,
and its value at the minimum of the loss function could be arbitrarily far away from
its own minimum. Figure 2.5 illustrates these problems for one data set. I therefore
concluded that both methods C and D are unusable for model selection.
25
150
20
100
15
10
50
5
200
400
600
800
200
400
600
800
Figure 2.5: Illustration of the problems encountered with regression method C.
The …gures show in phase space (N : prey, P : predator, IsoN : prey isocline, IsoP :
predator isocline) how the prey-dependent model has been …tted by method B (a,
process error …t) and by method C (b, weighted conditional least squares) to an
arti…cial stochastic time-series (created with a prey-dependent model). Method
C actually maximised dependence on initial conditions instead of minimising the
residuals, resulting in an obviously wrong …t. Similar problems were encountered
with method D. Note that the predator axes have different scales in (a) and (b).
A more interesting result is that only in 1% of all analysed time-series did methods
A and B both identify the wrong model. The most reliable model identi…cation is
therefore to apply several methods based on different assumptions about errors and
only accept results where all methods identi…ed the same model.
C. Jost
Identifying predator-prey models (PhD Thesis)
37
A side result from this analysis is the quality of parameters estimated by …tting
ODE’s to time-series data. I computed for each …t of a model to a time-series (that
was created by the same model) the ratio of the estimated parameter by the original
parameter (that created the time-series). The cumulative distributions of these ratios,
after …tting to the time-series with high error levels, are shown in Figure 2.6. If these
cumulative distribution curves pass through the point (1.0, 0.5) then the expected median is correct. The steeper these curves, the less the estimates are spread. We see that
only parameters r and K are reliably identi…ed in both models, all other parameters are
widely spread and the median is often far away from 1 (where it should be). All tested
error functions with this error level performed equally well (or bad).
The quality of the total parameter set (per model per time-series) can be visualised
in a similar way by working with the dominant eigenvalues (local stability) at the nontrivial equilibrium point. These eigenvalues have been computed algebraically to avoid
any numerical roundoff errors. Now there emerge strong differences between the error functions. Figure 7.5 shows the cumulative distribution functions of the ratios of
the dominant eigenvalues for all estimated parameter sets, error functions and models.
The steepness of each curve is approximately the same, therefore each error function
shows the same variation in the estimation of local stability, although there seems to
be a slightly smaller variation with ratio-dependent data. With respect to the second
criterion, deviation from the expected median, we see that observation error …t A performed globally best, followed by weighted process error …tting D. Method B, process
error …tting, always overestimated stability.
Conclusions from the ‘in silico’ approach
The result that should be retained from this analysis is that model identi…cation is
in principle possible with both studied error levels and for the time-series analysed
(dynamics reaching an equilibrium after one or two oscillations, starting from low initial
conditions and sampled at 20 times). Observation error …t is slightly more reliable
than process error …t, but the best identi…cation is obtained by …tting with both error
functions and only accepting model selections where both types identify the same model.
The estimated parameters are generally of poor quality and should be used with much
care. The method of …tting ordinary differential equations is therefore better suited for
model selection than for parameter estimation.
2.5.5
The ‘in vivo’ data analysis
As found in the previous section, model selection is most reliable when several goodnessof-…t criteria, based on different error models, select the same model. The two best
known error functions lead to observation error …t and process error …t. In consequence, I will only analyse real time-series that permit both types of …tting. This is
particularly restrictive for observation error …t, because the longer the time-series, the
less likely it is to get a reasonable …t of the whole trajectory. Of course, one could split
up long time-series into shorter pieces and …t trajectories to all pieces, but there is no
standard statistical method how this should be done correctly. To avoid any ambiguity
38
Identifying predator-prey models (PhD Thesis)
C. Jost
Prey-dependent data
1
1
1
r
0.6
K
1.4
1
Cumulative distribution
1
0.6
1.4
1
1
1.4
1
e/ c
0.6
1.4
1
1
h
0.6
a
µ
e
0.6
1
1.4
e/ c
0.6
1
1.4
e/ c
Ratio-dependent data
1
1
1
K
r
0.6
1.4
1
1
0.6
1
e/ c
1.4
0.6
1.4
1
1
1
h
0.6
1.4
1
e
0.6
1
e/ c
µ
1.4
0.6
1.4
1
e/ c
Figure 2.6: Analysis of the quality of each parameter estimate after …tting the
time-series with high error levels: cumulative distributions of the ratios e /c (e
is the estimated parameter, c the correct parameter that created the data) for all
error functions. — — —: observation error …tting A; ———: process error …tting
B; – – – –: weighted process error …tting (method D, see text).
linked to non-standard statistical methods I therefore only used time-series that can be
…tted reasonably to whole trajectories. These time-series range from simple protozoan
batch cultures (Gause 1935; Gause et al. 1936; Luckinbill 1973; Veilleux 1979; Flynn &
Davidson 1993; Wilhelm 1993), over spatially more complex laboratory predator-prey
systems (Huffaker 1958; Huffaker et al. 1963) to the very complex plankton systems
of freshwater lakes of the temperate zone (Carpenter et al. 1994; CIPEL 1995). The
plankton data from Carpenter et al. (1994) consist of edible phytoplankton and herbivorous zooplankton observed in two North American lakes from 1984 to 1990 (Paul
Lake and Tuesday Lake). Edible phytoplankton was de…ned by the authors to be all
C. Jost
Identifying predator-prey models (PhD Thesis)
39
phytoplankton with a biovolume less than that of a 30-m diameter sphere. In the
original data from Lake Geneva there were both edible phytoplankton (de…ned as organisms with length < 50m and biovolume below 104 m 3 ) and total phytoplankton, I
therefore …tted two systems: edible phytoplankton - herbivorous zooplankton, and total
phytoplankton - herbivorous zooplankton.
What are the differences between these real data and the arti…cial data analysed
in the previous section? The multiplicative nature of observation (and probably also
process) error has been con…rmed for all systems. However, the normality assumption
(for the log-transformed data) might be too strong, outliers (stronger tails, data points
farther away from the mean than expected) are possible in arthropod and in plankton
data. I therefore also …tted the data with an error function based on the sum of the
absolute values of the residuals in log scale (Laplacian or double exponential distribution). Another difference is the noise levels in the data that might be higher (especially
with phytoplankton) than the analysed noise levels. In this case, the 5% difference between the error functions after …tting both candidate systems (2.1) that was computed
in the previous section as assuring a 95% con…dence in the model selection might not
be enough. I therefore also performed a residual bootstrapping (as described by Efron
& Tibshirani (1993), see details in section 8.A) to compute a con…dence interval for
the error with each model and applying a standard t-test to see if one of the models
…ts signi…cantly better (with = 0.05). The same algorithm also yields an ‘improved
estimate of prediction error’ (IEPE, Efron & Tibshirani 1993) of the …tted model that
can be tested for signi…cance between both models. A …nal difference is that with real
data we do not know the process that created the data, that is, both candidate model
might be completely wrong. To cope with this problem I only considered …ts with a
likelihood above a threshold level that was estimated from the likelihoods obtained by
…tting to the arti…cial data. In sum, for each type of …tting (process error …t and observation error …t) four selection criteria were applied, and a model selection was only
considered signi…cant if all criteria (with sufficiently high likelihoods) with both error
functions yielded the same result.
The most signi…cant selection results are obtained with the protozoan data. Most
systems are either closer to prey dependence or the samples are too small to detect
predator dependence reliably. However, there is one predator-prey system with four
datasets (Flynn & Davidson 1993) that shows signi…cant predator dependence. The
predators in this system can show strong cannibalism at low prey-densities (personal
communication with the authors). Although such cannibalism was not observed in
the analysed data this suggests that the predators are capable of strong interference
when they encounter each other. It is also possible that heterogeneities developed in
the periods between the stirrings (every 12 h). Both factors might explain the highly
signi…cant support for the ratio-dependent model. To our knowledge this is the …rst
example of a protozoan system with monospeci…c prey and predator that shows this
strong predator dependence (compare to chapter 6). This exception illustrates that a
modeller has to know the biology of the system to be modelled, and that traits like
potential cannibalism can indicate that a model with predator dependence is more
appropriate. See section 8.5 for further comments and relations to other experiments.
There also seems to be no predator dependence in Huffaker’s arthropod data. In
40
Identifying predator-prey models (PhD Thesis)
C. Jost
most cases the prey-dependent model …ts better, and in both cases with process error
…t where the ratio-dependent model …ts better this …t is qualitatively wrong. Two
other aspects are important for the …ts to these data: 1) quantitatively the models …t
rather badly to the data, the experimental systems showing larger variation than can be
reproduced by our simple models and 2) trajectory …tting gives qualitatively correct …ts
with both models. The …rst point might be explained by Huffaker’s experimental setup,
food for prey being dispersed in a 2- or 3-dimensional structure and the prey colonising
this food in a fairly heterogeneous manner. Such a laboratory system is structurally
more complex than the protozoan batch cultures of the previous paragraph. The second
point indicates that the models used can nevertheless be used for qualitative analysis,
only quantitative conclusions should be interpreted with care.
The …ts to phyto- and zooplankton data are the most difficult to interpret. The
easiest conclusion would be that either the data are too noisy for this kind of model
identi…cation or that both models are too simple for lake dynamics. The …rst point is
supported by the qualitative nature of the process error …t regressions (mostly stable
or strongly stable systems) that might mean that the best prediction is not obtained by
dynamic nonlinear modelling but rather by simply using some mean value of the data as
a predictor. The second point is probably true for observation error …t, but not necessarily for process error …t. Despite these drawbacks, many signi…cant model identi…cations
are obtained with observation error …t, showing that there are long term dynamic patterns. These signi…cant …ts are of both types, prey- and ratio-dependent, with trends for
some lakes and for modelling edible phytoplankton only or total phytoplankton. However, these trends are not sufficiently convincing to give any recommendations when
which model might be more appropriate. Brett & Goldman (1997) argued that phytoplankton displays strong bottom-up in‡uence while zooplankton is more sensitive to
top-down control. The phytoplankton-zooplankton interaction itself (that is studied in
this thesis) is subject to both forces, which might also explain the ambiguity in model
identi…cation.
Conclusions from the ‘in vivo’ analysis
Systems with monospeci…c prey and monospeci…c predator in a homogeneous environment generally show little predator dependence and are better modelled with the
prey-dependent model. However, I found at least one system where the ratio-dependent
model …ts better than the prey-dependent model for all available time-series. Therefore,
before modelling such systems with a prey-dependent model, one should check whether
there are any biological traits that indicate strong predator dependence (e.g. potential
cannibalism). With lakes no conclusion in favour of one of the models can be drawn.
Whenever making predictions with models, such predictions should be cross-checked by
using another model. Prey dependence and ratio dependence offer promising frameworks to develop such alternative models and detect ‘robust’ predictions (robust in the
sense that they are independent of predator dependence) and direct further research if
no clear prediction is possible (e.g., to study how much predator dependence actually
occurs in a system or to parameterize an intermediate model).
C. Jost
Identifying predator-prey models (PhD Thesis)
41
De Mazancourt et al. (1998) and Zheng et al. (1997) are examples of studies that
have chosen this pluralistic modeling approach, both using on the one side a LotkaVolterra functional response (as an example of top-down control) and on the other side
a linear ratio-dependent functional response (to represent donor control). The …rst
study checked the robustness of some theoretical predictions, while the second study
contrasted these predictions to guide further research and the interpretation of …eld
results.
42
Identifying predator-prey models (PhD Thesis)
C. Jost
Chapter 3
Concluding remarks and
perspectives
3.1
General conclusions
The present thesis can be interpreted as a general validation of a ratio-dependent
predator-prey model. The mathematical properties of this model are thoroughly analysed, showing that global dynamics are stable coexistence of prey and predator, unstable coexistence (limit cycles), extinction of the predator only or extinction of both
prey and predator. For certain parameter values there are two basins of attraction, one
for the origin (extinction) and one for stable or unstable coexistence. This naturally
explains experimental results where coexistence or extinction were function of the initial
population levels. The model is shown to describe qualitatively correctly predator-prey
systems in freshwater plankton, protozoan batch and continuous cultures and laboratory
arthropod systems.
In a next step the ratio-dependent model is compared to an equally complex preydependent model, both qualitatively and quantitatively. The qualitative comparison
demonstrates that both models produce very similar dynamic patterns. The major
difference lies in the reaction to enrichment, which is destabilizing and increasing only
the predator equilibrium in the prey-dependent model, neutral with respect to stability
and increasing both prey and predator equilibrium in the ratio-dependent model. The
comparison of both models to the PEG-model (that summarises observed plankton
dynamics in freshwater lakes of the temperate zone and of different trophic states)
shows that they both can explain these dynamics if seasonality is added to one or more
parameters. The ratio-dependent model seems to explain the changes that occur in the
dynamics with enrichment slightly better than the prey-dependent model.
To compare the models quantitatively to predator-prey time-series I found the likelihood concept to be a …rm statistical framework. Protozoan data are generally better
described by the prey-dependent model. However, there is also one protozoan system
where the ratio-dependent model describes the dynamics more accurately. The analysed
arthropod data correspond better to the prey-dependent model. For the phytoplanktonzooplankton interaction both models are valid and none of them better than the other.
43
44
Identifying predator-prey models (PhD Thesis)
C. Jost
Consequently, predictions for freshwater plankton systems should be based on both
models to avoid falling victim to mathematical artifacts of one of them (such as the
paradox of enrichment).
There exist for both the prey-dependent and the ratio-dependent model equivalent
forms in microbiology (Monod’s and Contois’ model). Microbiologists have also done
the kind of quantitative and qualitative comparison that is applied in this thesis. Reviewing their results I found that Monod’s model seems correct with monospeci…c prey
and monospeci…c predators in a homogeneous environment. However, every deviation
from these conditions induces predator dependence in the growth function (numerical
response), such that often Contois’ model describes the observed dynamics better than
Monod’s model. The main …eld of application of Contois’ model is in plurispeci…c and
heterogeneous systems such as fermentation processes and waste water treatment.
In conclusion, the prey-dependent model seems most appropriate in cases of monospeci…c prey and monospeci…c predators in homogeneous environments. In all other cases
any degree of predator dependence in the functional response can be observed and the
best suited model is a priori unknown. Prey-dependent and ratio-dependent models
can be interpreted as two extrema with respect to predator dependence. Using both
helps detecting predictions that are sensitive to predator dependence and direct further
research if necessary.
3.2
Perspectives for continuation of work
In the following I will list some possible lines of research that could further enhance the
work presented in this thesis.
3.2.1
A non-parametric approach
The technique used for model selection in chapters 7 and 8 requires to parameterize the
whole predator-prey model, in particular also the prey growth function and the predator
mortality function which are of no interest to the detection of predator dependence in
the functional response. Even worse, inaccuracies in the chosen formulations (logistic
growth and constant predator mortality rate in the case of this thesis) could in‡uence the
detection of predator dependence. Another possible problem is the absence of delayed
effects in the predator-prey model. A delay in the numerical response with respect
to the functional response is theoretically plausible and has improved the goodnessof-…t in Harrison’s (1995) reanalysis of Luckinbill’s (1973) protozoan data and in the
bacteria-bacteriophage system of Bohannan & Lenski (1997).
An elegant solution for both problems has been outlined by Ellner et al. (1997).
The authors study the system
dx(t)
= f (x(t )) + g(x(t)).
dt
Based on a time-series approach they develop a method to estimate the delay parameter
and to reconstruct f and g non-parametrically. Tested with arti…cial time-series,
C. Jost
Identifying predator-prey models (PhD Thesis)
45
adding both observation and process error with CV = 0.1, they …nd that a correct
estimation of is possible for time-series as short as 100 data points. The non-parametric
reconstruction of f and g looks fairly well and the authors point out that their method
can reveal information about the biology of f and g.
This is exactly what could be done to identify the functional response. Estimating
the delay in the numerical response (as a byproduct) and reconstructing the functional
response non-parametrically with con…dence intervals on all involved variables one could
now …t the functional response directly to the reconstructed data (also using errors-invariables techniques) and identify empirically the most parsimonious model to describe
it. Protozoan systems such as studied in Veilleux (1979) could provide the data necessary
for such an analysis.
3.2.2
Paying more attention to deterministic extinction
The quasi extinction risk of a natural population is usually estimated by parametrising
a model of the population (e.g., with a Leslie model, Leslie 1948), doing stochastic
simulations and measuring the proportion of simulations during which the population
abundance fell below a prede…ned threshold value. One of the reasons for this approach
is that typical population models do not allow for deterministic extinction. Working
also with models that include this feature (such as general ratio-dependent food webs or
food webs extended from the ideas in the predator-prey models of section 2.3.1) could
provide additional information on the risks of endangered populations. Extinction as a
deterministic outcome has not been considered at all in the ecological literature.
3.2.3
Modelling plankton data more realistically
I analysed the plankton data from Lake Geneva only in the predator-prey context with
constant exogenous factors (in form of model parameters). Actually, there are much
more data available: plankton abundances for each species, temperature at several
depths of the Lake, nitrogen and phosphorous content in the water. Using these external variables as model input a much better prediction of total phyto- and zooplankton
abundances one or two weeks ahead might be possible. However, while adding inputs
will always increase goodness-of-…t, this also increases model complexity (with all its
inconveniences: more parameters need being estimated, it becomes more difficult to
transfer the model to another ecosystem, etc.) and the predictions may not become
signi…cantly better by adding these inputs (likelihood ratio tests, information criteria).
The identi…cation of the most important driving external variables is an old problem in
ecology. The usual statistical approach uses prinicpal component analyses. However,
PCA only detects linear dependencies, nonlinearities are only detected as noise. Empirical …tting methods such as neural networks also detect nonlinear patterns (Lek et al.
1996). By splitting the data set in two (in-sample and out-of-sample) one can teach the
neural network on the in-sample data and test its predictive power on the out-of-sample
data. This procedure can be combined with bootstrap techniques. Such methods could
detect the two, three or four most important external variables also with respect to
46
Identifying predator-prey models (PhD Thesis)
C. Jost
nonlinear dependencies. These variables could then be used to develop a simple explicit
dynamic model to predict the variables of interest, e.g., total phyto- and zooplankton.
3.2.4
How to treat process and observation error together?
The statistical methods I used in this thesis all make the assumption that there is only
one type of error in the data, observation error or process error. This is in contrast to the
data which always contain both types of stochasticities. Statisticians have developed
several strategies how to cope with this problem. In this section I will discuss three
possible strategies. All of them are computationally more intensive than the techniques
I used, but with the next generation of personal computers the actual computing time
might decrease to acceptable values.
Hilborn & Walters (1992) propose to …t the models by assuming that all errors are
of one type only (e.g. observation error) and to assess a possible bias by stochastic
simulation. In the context of continuous models this means …tting data produced by a
stochastic ordinary differential equation (SODE) to a deterministic ordinary differential
equation (ODE). This approach may be well suited for parameter estimation if there
is no doubt about the model structure. However, we do no not know whether the expected trajectory of a SODE maintains the structure of the underlying ODE (which is
important for a correct model selection). This is the case for single population exponential growth (Yodzis 1989) and single population logistic growth (Nisbet & Gurney
1982; Braumann 1983), but these results rely on the existence of an analytic solution
of the ODE and I am not aware of any such work for more complex systems. Some
preliminary studies show that there may be a difference between expectation and the
deterministic model: I simulated SODE’s by adding white noise with a standard deviation proportional to the actual population size (thus simulating the characteristics of
a lognormal process error, see also the arti…cial data creation in chapter 7), computed
a large number of replicates and calculated from them the mean population dynamics (Figure 3.1, dotted curves). Fitting the deterministic model (whole trajectory) to
these expectations showed that often the trajectories could not be approximated by the
deterministic ODE, and model identi…cation became very difficult. Figure 3.1 shows
moderate examples of these deviations from deterministic behaviour, together with the
best …t obtainable to the deterministic model that was used as the core of the SODE
and the deterministic trajectory with the original parameters. These two examples illustrate that the structure of the expectation of a SODE can be different from the one
of its deterministic core function. Further numerical work would be needed to elaborate
how stochastic noise can change the actual structure of a SODE.
Errors-in-variables (having uncertainty in both the dependent and the independent
variable) offers another approach how time-series could be …tted to models. The idea is
to estimate not only the parameters, but also the actual values of the state variables for
each measurement (the so-called nuisance parameters). This can be done in a likelihood
framework, but requires independent knowledge of one of the errors. In their milestone
paper Reilly & Patino-Leal (1981) describe such estimation techniques for models that
are linear or nonlinear in the parameters. They propose nested iterations to estimate
alternatingly parameters and state variables. Schnute (1994) and Schnute & Richards
C. Jost
Identifying predator-prey models (PhD Thesis)
prey-dependent data
47
ratio-dependent data
(a)
abundances
(b)
time
time
Figure 3.1: Two examples illustrating the problem that expectations of SODE’s
can have a different structure than the deterministic core function of the SODE.
Diamonds and stars are prey and predator data (respectively) of the series of
expectations (obtained by Monte Carlo simulation), the continuous line is the
deterministic trajectory with the original parameters and the dashed line shows
the best …t of the deterministic model to the expected trajectory. (a) is an example
for the prey-dependent model where the expectation strongly deviates from the
original trajectory while it can still be …tted reasonably to the deterministic core
model. (b) shows an example for the ratio-dependent model where the expectation
is closer to the original trajectory but it cannot be …tted to the deterministic model.
(1995) discuss the errors-in-variables technique in the context of …sheries management.
In other branches of ecology the method does not seem to be used actively.
Yet another set of approaches are simulation-based regression techniques (Ellner &
Turchin 1995; Gouriéroux & Monfort 1996). The idea is to use the stochastic model to
simulate a reference data set which is then used to estimate the discrepancy between the
real data and this reference data set (inspired from reconstruction techniques used in
chaos theory, see Kantz & Schreiber 1997 for a very readable introduction). The parameters in the stochastic model are chosen such that the discrepancy becomes minimal.
As in the errors-in-variables method this technique requires independent knowledge of
one of the errors. Little is known about the reliability of parameter estimation and
model selection when applying this method to typical ecological time-series.
Partie If
Introduction générale et résultats
principaux
1f
Chapitre 1f
Prédateur-dépendance de la
réponse fonctionnelle et
implications en écologie
1.1f
Quelques commentaires concernant les modèles
mathématiques en écologie des populations
Une des premières descriptions de la dynamique d’une population se trouve vers la …n
du 18ème siècle quand Malthus (1798) introduisit ce qui est connu aujourd’hui sous le
nom de ‘croissance Malthusienne’ : une population augmente exponentiellement tant
que ses ressources ne sont pas limitantes. En désignant par N (t) l’abondance de la
population au temps t cette croissance peut être décrite par l’équation différentielle
dN (t)
= rN (t)
dt
où r est le taux de croissance. Mais que se passe-t-il si les ressources (par ex. l’espace,
la nourriture, les nutriments essentiels) sont limitées? L’idée qu’une ressource limitante
peut arrêter la croissance d’une population a été introduite empiriquement par Verhulst
(1838) dans ce qui est appelé aujourd’hui le modèle de croissance logistique,
dN
= rN (1
dt
N
).
K
(1f.1)
La capacité de soutien K désigne une abondance limite au-dessus de laquelle la croissance de la population devient négative, tandis qu’au-dessous, la croissance est positive.
K représente donc une valeur d’équilibre vers laquelle l’abondance de la population converge. K peut aussi être interprétée comme une mesure des ressources disponibles. Le
modèle logistique est donc un modèle de contrôle de bas en haut (donor control, Pimm
1982) : la population est contrôlée par ses ressources.
En fait, il existe un autre mécanisme qui empêche la population de croître exponentiellement : le cas où elle est consommée par une autre population à un taux excédant
son taux de croissance. Une telle interaction proie-prédateur a été décrite originellement
3f
4f
Identi…cation de modèles proie-prédateur (thèse de doctorat)
C. Jost
par deux chercheurs travaillant indépendamment, Lotka (1924) et Volterra (1926). En
désignant par P (t) l’abondance de cette deuxième population, le prédateur, l’interaction
est décrite par le jeu d’équations différentielles
dN
dt
dP
dt
= rN
aNP
= eaNP
P
(1f.2)
où a représente le taux d’attaque, e l’efficacité de conversion (pourcentage de la biomasse consommée qui est convertie en biomasse de prédateur), et le taux de mortalité
du prédateur. Ce système va faire des cycles éternels qui passent périodiquement par
les valeurs initiales des abondances de la proie et du prédateur (stabilité neutre). La
découverte d’un système mathématique proie-prédateur coïncidait avec l’article d’Elton (1924) sur les ‡uctuations des populations animales, ce qui a été à l’origine de
l’hypothèse selon laquelle ces ‡uctuations résultent d’interactions entre proies et prédateurs. Des centaines d’articles ont examiné cette hypothèse, mais il n’y a, encore
aujourd’hui, que peu de réponses claires (revue dans Chitty 1996). Cependant, les
populations ‡uctuantes ne sont pas le sujet de cette thèse.
Rappelons que le contrôle des abondances est effectué de bas en haut (donor control)
dans le modèle logistique. Au contraire, dans le modèle proie-prédateur de LotkaVolterra, l’équilibre de la proie s’écrit
N? =
.
ea
Donc, non seulement le prédateur empêche la proie de croître indé…niment, mais de
plus ses paramètres , e et a dé…nissent l’abondance d’équilibre de la proie. Cette
relation exprime mathématiquement ce qui est souvent appelé contrôle de haut en bas
(top-down).
Il n’existe pas en écologie de dé…nition unique et consensuelle des termes de contrôle
(ou effet) s’exerçant ‘de bas en haut’ ou ‘de haut en bas’, en particulier dans les modèles
de chaînes trophiques. En général, le terme de contrôle par le haut est utilisé dans le
sens de Hairston et al. (1960), c’est à dire pour exprimer l’idée selon laquelle chaque
niveau trophique est potentiellement capable de contrôler l’abondance de sa proie en la
maintenant à un niveau bas et indépendant de l’abondance des ressources de la proie.
Tout niveau trophique exerce ce contrôle sur sa proie, à moins qu’il ne soit lui-même
contrôlé par son propre prédateur. Ce modèle explique que ‘le monde est vert’ par le
fait que, dans une chaîne à trois niveaux trophiques, les herbivores sont contrôlés par
les carnivores et n’empêchent donc pas les plantes de croître selon l’abondance de leurs
ressources. Une généralisation de cette théorie est l’hypothèse ‘EEH’ (Exploitation
Ecosystem Hypothesis Fretwell 1977; Oksanen et al. 1981) selon laquelle le niveau
trophique le plus haut d’une chaîne, ainsi que ceux qui en sont éloignés par un nombre
impair de niveaux, sont contrôlés par le haut, tandis que les autres sont contrôlés par le
bas. Un cas particulier de cette théorie est celui de la cascade trophique (Carpenter et al.
1985; Carpenter & Kitchell 1994) qui examine la façon dont des changements en haut de
la chaîne trophique se répercutent plusieurs niveaux trophiques en dessous. Les effets
s’exerçant de bas en haut sont prépondérants dans les modèles de chaîne trophiques
C. Jost
Identi…cation de modèles proie-prédateur (thèse de doctorat)
5f
dites ‘donor-controlled’ (contrôlées par le donneur) (Pimm 1982; Berryman et al. 1995)
dans lesquels l’abondance au sein de chaque niveau trophique dépend uniquement de ses
ressources, et où les taux de mortalité sont indépendants de l’abondance des prédateurs.
Pour plus d’informations, voir les discussions dans Hunter & Price (1992), Strong (1992),
Power (1992) et Menge (1992) ou la synthèse, très bien formulée, du chapitre II de
Ponsard (1998). Dans la présente thèse, je parlerai de contrôle par le bas lorsque
l’abondance d’équilibre d’un niveau trophique est corrélée positivement avec l’abondance
des ressources du premier niveau trophique dans le système étudié (par ex. K dans le
cas d’une croissance logistique). Les populations seront dites contrôlées par le haut
dans le cas où une modi…cation de l’abondance ou des paramètres caractéristiques de
ses prédateurs induit un changement d’abondance à l’équilibre de cette population.
Comme nous le verrons par la suite, ces dé…nitions ne s’excluent pas mutuellement.
Une population peut être contrôlée à la fois par le bas et par le haut.
Les deux théories - contrôle par le haut ou par le bas - font des prédictions différentes
sur les réponses des abondances d’équilibre de différents niveaux trophiques à une augmentation des ressources (voir Tableau 1f.1). Par exemple, pour réduire l’abondance
des algues dans un lac à quatre niveaux trophiques (algues, zooplancton herbivore, zooplancton carnivore, poissons planctonophages), la théorie du contrôle par le bas dit qu’il
faut réduire l’abondance du lac en nutriments, tandis que la théorie du contrôle par le
haut (Oksanen et al. 1981) suggère au contraire que c’est une fertilisation du lac qui
devrait permettre de réduire l’abondance des algues.
Revenons maintenant au système de Lotka-Volterra : tous les systèmes proie-prédateur
sont-ils contrôlés par le haut, comme le suppose ce modèle? Où existe-t-il d’autres systèmes proie-prédateur correspondant plutôt à un contrôle par le bas ou à un contrôle
mixte? Ces questions sont d’un intérêt général car toutes les interactions directes entre
espèces de différents niveaux trophiques sont du type proie-prédateur. Le choix d’une
forme mathématique particulière pour décrire cette interaction proie-prédateur dans les
réseaux trophiques en général peut avoir des effets importants sur les prédictions de ces
modèles quant aux réactions des écosystèmes à des perturbations en bas (changement du
statut trophique) ou en haut de la chaîne (introduction ou élimination d’un prédateur).
Ces prédictions sont utilisées pour décider des politiques de gestion des populations naturelles exploitées, ou en biologie de la conservation. L’importance relative des forces de
haut en bas et de bas en haut dans les systèmes naturels est mal connue (Menge 1992;
Power 1992). Dans une telle situation, un écologiste devrait soit faire des prédictions
qui sont indépendantes du type de contrôle, soit identi…er les prédictions qui sont sensibles à la dominance de l’un ou l’autre type de contrôle. Je vais introduire ci-dessous un
modèle proie-prédateur avec des caractéristiques du contrôle de bas en haut, modèle qui
peut servir d’alternative aux modèles du type Lotka-Volterra. En confrontant, de façon
qualitative et quantitative, ce modèle ainsi que des modèles du type Lotka-Volterra avec
des séries chronologiques, je cherche à obtenir des informations sur les mécanismes de
contrôle à l’oeuvre dans ces systèmes.
6f
1.1.1f
Identi…cation de modèles proie-prédateur (thèse de doctorat)
C. Jost
Un cadre général pour la modélisation de systèmes proieprédateur
Les modèles proie-prédateur qui vont être étudiés dans cette thèse suivent deux principes.
Le premier est que la dynamique de la population peut être décomposée en processus
de naissance et processus de mort,
dN
= croissance
dt
mortalité .
Le deuxième principe est la conservation de la masse (Ginzburg 1998), qui dit que la
croissance du prédateur est une fonction directe de ce qu’il a mangé. Avec ces deux
principes, on peut écrire la forme canonique d’un modèle proie-prédateur de la façon
suivante :
dN
dt
dP
dt
= f (N )N
g(N, P )P
= eg (N, P )P
N (N )N
P (P )P,
(1f.3)
où f (N ) représente la croissance de la proie, g(N, P ) la réponse fonctionnelle (proies
mangées par prédateur par unité de temps, Solomon 1949), et N (N ) et P (P ) la
mortalité naturelle, respectivement de la proie et du prédateur. La réponse numérique
(taux de croissance du prédateur en fonction de la consommation) est considérée comme
étant proportionnelle à la réponse fonctionnelle. Il faut noter que le système (1f.3)
devient le système de Lotka-Volterra lorsque f (N ) = r, g(N, P ) = aN , N (N ) = 0 et
P (P ) = . Beaucoup d’autres formes pour les fonctions f , g et x (x ∈ {N, P }) ont été
proposées dans la littérature (voir Appendice A pour une collection (non exhaustive)
de formes fonctionnelles que j’ai trouvées dans la littérature). C’est en particulier la
réponse fonctionnelle g qui a initié un grand nombre de travaux (Bastin & Dochain
(1990)), avec une collection de plus de 50 modèles.
On fait habituellement l’hypothèse que la consommation est la cause principale de
mortalité de la proie, de sorte que N (N ) peut être négligé (tant que le prédateur existe).
Je suivrai cette hypothèse dans toute cette thèse. Je remplacerai également f (N ) par
la croissance logistique standard (Verhulst 1838, voir Figure 2f.3) et considérerai que la
mortalité du prédateur est constante, comme dans les équations de Lotka-Volterra. Ces
hypothèses donnent le système
dN
dt
dP
dt
= r 1
N
K
= eg (N, P )P
N
g(N, P )P
P.
(1f.4)
C. Jost
1.1.2f
Identi…cation de modèles proie-prédateur (thèse de doctorat)
7f
Le contrôle de haut en bas et de bas en haut dans les
modèles proie-prédateur
Traditionnellement, on fait l’hypothèse que la réponse fonctionnelle dans le système
(1f.4) ne dépend que de la proie, g(N, P ) = g(N ), par ex.
g(N ) = aN
aN
g(N ) =
1 + ahN
(Lotka 1924 )
(Holling 1959b )
(1f.5)
(1f.6)
(voir tableau A.2 pour d’autres exemples). Je quali…erai ce type de réponse fonctionnelle
proie-dépendante. Dans ce cas, l’équilibre de la proie est de la forme
?
1 N =g
.
e
On voit que, comme dans le système de Lotka-Volterra, cet équilibre est entièrement
dé…ni par les paramètres du prédateur. Graphiquement, ceci se voit dans l’isocline
vertical du prédateur dans le graphique 4.2a. Une augmentation de la capacité de
charge K (enrichissement) va changer seulement l’isocline de la proie et va entraîner
une déstabilisation de l’équilibre. Ce phénomène est connu sous le nom de ‘paradoxe de
l’enrichissement’ (Rosenzweig 1971). L’approche traditionnelle proie-dépendante dans
le cadre du système (1f.4) va donc toujours aboutir à un modèle proie-prédateur contrôlé
de haut en bas avec un effet déstabilisateur de l’enrichissement.
Cette analyse indique la façon dont il faut modi…er le système (1f.4) pour changer
ce pattern : la réponse fonctionnelle doit aussi dépendre de l’abondance du prédateur,
g = g(N, P ). J’appellerai ce cas prédateur-dépendance. En principe, une densité elevée
de prédateurs entraîne un nombre accru de rencontres entre prédateurs. Ceci peut
entraîner une diminution dans l’efficacité de la prédation, soit seulement parce que
l’organisme perd du temps pour détecter que l’autre individu est aussi un prédateur
(et pendant ce temps le prédateur ne cherche pas d’autres proies), soit à la suite d’une
interférence active entre les deux prédateurs (Beddington 1975; Hassell & Varley 1969).
L’abondance du prédateur devrait par conséquent avoir un effet négatif sur la réponse
fonctionnelle,
dg (N, P )
< 0.
dP
Dans la prochaine section, nous discuterons plus en détail la façon dont cette modi…cation affecte le pattern du contrôle de haut en bas. Il existe une vaste littérature sur
l’interférence entre prédateurs (voir les revues dans Hassell (1978a) et dans Sutherland
(1996)).
En éco-éthologie, on trouve aussi des exemples de l’effet inverse : la facilitation
du prédateur, où une coopération entre les prédateurs peut avoir pour effet un succès
reproducteur élevé. Des exemples en sont les lions chassant en paires plutôt que seuls
(Krebs & Davies 1993, p. 278), ou des coquillages marins (patelles) qui capturent
des frondes d’algues et des débris de macrophytes marins (Bustamante et al. 1995).
Cependant, dans cette thèse je ne considérerai que le cas d’une in‡uence négative des
prédateurs sur la réponse fonctionnelle, ce qui semble être plus fréquent dans la nature.
8f
1.2f
Identi…cation de modèles proie-prédateur (thèse de doctorat)
C. Jost
Introduction de la prédateur-dépendance
La prédateur-dépendance dans la réponse fonctionnelle a été introduite à plusieurs reprises. La première introduction avec un certain écho dans la littérature a été faite par
Hassell & Varley (1969) , qui proposaient que le taux d’attaque a devrait diminuer avec
une augmentation de la densité des prédateurs. Les auteurs ont proposé une décroissance
exponentielle avec un taux m,
a = P
m
.
(1f.7)
Le paramètre m peu être interprété comme un coefficient d’interférence. Plus les prédateurs interfèrent entre eux, plus m est grand. Au début ce concept n’était utilisé
que pour expliquer les taux d’attaque observés sur le terrain. L’incorporation de ce
concept dans le système (1f.4) a aussi démontré un effet stabilisant (Beddington 1975;
Beddington et al. 1978). Il y a deux problèmes techniques avec cette formulation. Le
premier problème concerne les dimensions (Beddington et al. 1978), et peut être corrigé
en introduisant un paramètre muet uP ayant la dimension dimension [P ] (qui représente
une ‘unité’ de prédateur) et en reécrivant (1f.7) sous la forme
m
P
a=
.
uP
Habituellement, ce paramètre muet n’est pas mentionné mais supposé implicitement
(Hassell 1978a, p. 84). Le deuxième problème est plus grave et concerne l’analyse
mathématique de systèmes qui utilisent ce type de taux d’attaque : en raison de la
forme exponentielle de m, il est souvent impossible de trouver des équilibres explicites.
Ceci rend l’analyse de stabilité et d’autres types d’analyse plus difficiles car nécessitant
la théorie des fonctions implicites ou d’autres techniques sophistiquées.
Ces deux problèmes peuvent être évités en introduisant la prédateur-dépendance de
la façon proposée par Beddington (1975) et DeAngelis et al. (1975),
g(N, P ) =
aN
1 + ahN + cP
(1f.8)
où h est le temps de manipulation (le même que celui utilisé dans la réponse fonctionnelle
classique Holling type II (1f.6)) et c est une constante empirique. Beddington (1975) a
dérivé ce modèle basé sur des mécanismes comportementaux, faisant l’hypothèse que le
prédateur passe son temps à trois activités : chercher des proies, manger des proies ou
manipuler d’autres prédateurs (reconnaissance, éventuellement interférence). c devient
dans ce contexte le produit du taux de rencontre avec d’autres prédateurs et du temps
de manipulation d’un autre prédateur.
Le système (1f.4) avec cette réponse fonctionnelle peut être analysé par les techniques
traditionnelles, en déterminant les équilibres et en analysant leur stabilité (par ex. avec
le critère de Routh-Hurwitz, Edelstein-Keshet 1988). C’est ainsi que DeAngelis et al.
(1975) se sont aperçus que, dans le cas où 1 + ahN cP , ce système devient contrôlé
de bas en haut. La réponse fonctionnelle (1f.8) fournit donc une possibilité simple
d’incorporer des traits typiques pour le contrôle de bas en haut. Cependant, il est
C. Jost
Identi…cation de modèles proie-prédateur (thèse de doctorat)
9f
rarement utilisé, que ce soit dans des études mathématiques ou dans des applications à
des systèmes réels. Plusieurs raisons peuvent expliquer ce fait : d’une part, le paramètre
c est difficile à estimer, et d’autre part l’analyse mathématique est moins simple que,
par exemple, dans des systèmes où la réponse fonctionnelle est du type Lotka-Volterra
ou Holling type II, car cette réponse fonctionnelle est fonction de deux variables, N et
P , d’où la nécessité du recours à la différentiation partielle.
1.2.1f
Une approche nouvelle : la réponse fonctionnelle ratiodépendante
Une manière plus simple d’incorporer la prédateur-dépendance a été proposée par Arditi
& Ginzburg (1989), en modélisant la réponse fonctionnelle comme fonction d’un seul argument (comme dans l’approche traditionnelle), cet argument étant toutefois le rapport
entre l’abondance des proies et celle des prédateurs,
N
(1f.9)
g(N, P ) = g
.
P
Ceci est un cas particulier de prédateur-dépendance. Les modèles comportant ce type
de réponse fonctionnelle sont quali…és de ratio-dépendants (Arditi & Ginzburg 1989).
g(x) est en général une fonction croissante, continue et bornée de son argument x,
similaire à la fonction proie-dépendante du type Holling II. Une version particulière
fréquemment utilisée du modèle ratio-dépendant est fondée sur une analogie avec cette
réponse fonctionnelle Holling II (1f.6),
N
N/P
(1f.10)
g
=
.
P
1 + hN/P
La simplicité de cette approche a permis à Arditi & Ginzburg (1989) d’analyser
des chaînes trophiques dans le cas général, et de comparer les réponses des abondance
d’équilibre des différents niveaux trophiques à un changement de la productivité primaire (F := f (N )N ). Le résultat est résumé dans le tableau 1f.1. On voit que, dans des
chaînes ratio-dépendantes, les équilibres de tous les niveaux trophiques augmentent avec
la productivité primaire (indépendamment de la longueur de la chaîne), tandis que, dans
des chaînes proie-dépendantes, on observe une alternance d’augmentations et d’absences
de changement ou de diminutions, selon de la distance au niveau trophique le plus élevé.
En particulier, l’avant-dernier niveau est toujours contrôlé de haut en bas, tandis que
les niveaux trophiques qui se trouvent un nombre pair de niveaux sous le niveau le plus
élevé sont toujours contrôlés de bas en haut. Le pattern du modèle proie-dépendant a
été observé en laboratoire dans des chaînes trophiques à trois niveaux (Kaunzinger &
Morin 1998). Cependant, une légère augmentation (non expliquée) du deuxième niveau
trophique avec un enrichissement a été observée dans cette étude. Oksanen et al. (1981),
en analysant des chaînes trophiques dans le cas général, pensaient avoir trouvé le pattern
prédit par la théorie proie-dépendante dans des données provenant des communautés
végétales de la zone arctique et subarctique (Oksanen 1983). Les meilleures données
proviennent de lacs d’eaux douces le long d’un gradient d’abondance de nutriments.
10f
Identi…cation de modèles proie-prédateur (thèse de doctorat)
C. Jost
Table 1f.1: Tendances des équilibres de la ressource basale (N ? ) et des niveaux
trophiques de prédateurs (P1? , P2? , etc.) lorsque le production primaire F augmente, dans des chaînes proie-dépendantes et ratio-dépendantes. Les ‡èches indiquent le changement prédit. Tableau adapté de Arditi & Ginzburg (1989).
Proie-dépendance
Ratio-dépendance
Niveau trophique
2 3
4
5
2
3
4
5
?
P4
%
%
?
P3
% →
% %
P2?
% → %
% % %
P1?
% → % &
% % % %
N?
→ % & %
% % % %
Des études extensives ont montré que la biomasse de tous les niveaux trophiques (phytoplancton, zooplancton herbivore, zooplancton carnivore et poissons planctivores) augmentent le long d’un gradient d’enrichissement (McCauley et al. 1988; Mazumder 1994;
Mazumder & Lean 1994; McCarthy et al. 1995; Harris 1996). Les observations concordent donc qualitativement avec les prédictions du modèle ratio-dépendant. Plusieurs
explications ont été avancées pour expliquer ce pattern (McCauley et al. 1988; Abrams
1993). La prédateur-dépendance dans la réponse fonctionnelle n’est qu’une possibilité. Cependant, la ratio dépendance est l’une des explications les plus parcimonieuses.
Notons également que l’utilisation d’une réponse fonctionnelle ratio-dépendante dans
le système (1f.4) in‡uence le lien entre stabilité et enrichissement (voir chapitre 5 pour
plus de détails) : un changement de K n’in‡uence que l’amplitude des trajectoires, mais
pas la stabilité locale de l’équilibre. Ceci signi…e que le paradoxe de l’enrichissement est
fortement lié à l’isocline verticale du modèle proie-dépendant, et que toute variation de
cette isocline peut permettre d’autres réponses qu’une déstabilisation.
La modélisation de réseaux trophiques entiers plutôt que de chaînes trophiques est
mathématiquement bien plus complexe. Michalski & Arditi (1995a) et Arditi & Michalski (1995) ont développé des réseaux trophiques mathématiques avec des réponses fonctionnelle du type Holling II, DeAngelis-Beddington et ratio-dépendantes. Avec cette
dernière, un trait distinctif s’est révélé qui n’existe pas avec les autres réponses fonctionnelles : lorsque les dynamiques sont loin de l’équilibre, des liens trophiques apparaissent
et disparaissent fréquemment (Michalski & Arditi 1995b). Une fois à l’équilibre, il n’y
a que peu de liens dans le réseau, mais si l’on cumule des observations faites sur le
réseau pendant différentes phases transitoires de la dynamique, on obtient un réseau
trophique d’une apparence très complexe tel que celui de la morue dans Lavigne (1996)
(reproduction dans Yodzis (1995)).
1.2.2f
Données expérimentales en faveur de la ratio-dépendance
En dehors des corrélations entre abondances de plancton et l’état trophique des lacs,
citées plus haut, il existe également des résultats expérimentaux en faveur de la ratiodépendance. Une série d’expériences ont été effectuées sur des cladocères herbivores,
Daphnia magna et Simocephalus vetulus, consommateurs de l’algue Chlorella vulgaris
C. Jost
Identi…cation de modèles proie-prédateur (thèse de doctorat)
11f
(Arditi et al. 1991b; Arditi & Saïah 1992). Ces algues ont été cultivées dans des compartiments isolés et ajoutées à taux constant à un premier compartiment contenant
l’un ou l’autre prédateur. Les effluents de ce compartiment (ne contenant que des
algues, les prédateurs étant empêchés de sortir) étaient transférés dans un deuxième
compartiment contenant des prédateurs etc. (voir …gure 1f.1). L’hypothèse était que,
dans le cas d’une isocline verticale du prédateur (proie-dépendance), le prédateur ne
devrait subsister que dans le premier compartiment, broutant les algues à un taux …xe
(dé…ni par les prédateurs) ne permettant pas l’existence de prédateurs dans les compartiments suivants. Au contraire, si l’isocline des prédateurs est inclinée, ceux-ci peuvent
théoriquement survivre dans tous les compartiments (voir …gure 1f.1). Les prédateurs
qui ont été choisis pour l’expérience ont un comportement spatial particulier, Daphnia nageant d’une manière homogène, tandis que Simocephalus se tient plus près des
bords du compartiment. Daphnia a suivi le pattern prédit par l’isocline verticale, tandis que Simocephalus s’est maintenu dans plusieurs compartiments consécutifs. Dans
une deuxième étape, Daphnia a été empêchée de nager dans le compartiment total,
tandis que Simocephalus a été distribué d’une manière homogène par agitation du récipient. Avec ces modi…cations, Daphnia persistait dans plusieurs compartiments et
Simocephalus seulement dans le premier. Ces expériences démontrent clairement que
l’hétérogénité spatiale peut provoquer une isocline du prédateur inclinée. Néanmoins,
ils ne démontrent pas que cette isocline passe par l’origine, comme le prédit le modèle
ratio-dépendant (Ruxton & Gurney 1992; Holmgren et al. 1996). La Figure 1f.1 illustre ceci pour l’exemple de trois compartiments consécutifs. Le modèle ratio-dépendant
et le modèle de DeAngelis-Beddington prédisent qualitativement le même pattern, en
particulier le fait que l’abondance du prédateur à l’équilibre décroît géométriquement
d’un compartiment à l’autre.
Une autre étude (Arditi & Akçakaya 1990) a consisté en une réanalyse de données
expérimentales sur des systèmes proie-prédateur-hôte parasitoïde. Dans ces expériences,
on a fait varier les abondances de proies et de prédateurs, et la réponse fonctionnelle a
été mesurée. Arditi & Akçakaya ont ajusté un modèle fondé sur la fonction de Hassell
& Varley (1969) (1f.7),
N
P m N
g(N, P ) = g
.
=
(1f.11)
Pm
1 + P m hN
Notons que cette réponse fonctionnelle, nommée Hassell-Varley-Holling, est proie-dépendante pour m = 0, tandis qu’elle est ratio-dépendante pour m = 1. En fait, l’estimation
elle-même a été faite en deux étapes, en estimant tout d’abord le taux d’attaque a =
P m pour chaque densité de prédateur par régression habituelle du modèle de Holling
type II (régression non-linéaire, tenant compte de la diminution des proies pendant
l’expérience grâce à l’utilisation de l’équation du prédateur aléatoire, Rogers 1972),
g (N | P ) =
aN
,
1 + ahN
et en estimant ensuite le paramètre m par régression linéaire de la relation (1f.7) en
coordonnées logarithmiques. Les estimations de m venant de cette procédure étaient
majoritairement plus proche de 1 que de 0, indiquant non seulement que l’isocline du
12f
Identi…cation de modèles proie-prédateur (thèse de doctorat)
C. Jost
Figure 1f.1: Prédictions de l’abondance du prédateur à l’équilibre dans l’expérience
en compartiments successifs décrite dans Arditi et al. (1991b), pour un modèle
ratio-dépendant (à gauche) et un modèle du type DeAngelis-Beddington (à droite).
N est la concentration d’algues et P le nombre de cladocères. Ison et Isop sont
respectivement l’isocline de la proie et l’isocline du prédateur. D représente un
cladocère individuel. Le premier compartiment s’équilibre comme indiqué en (a)
et (b), laissant passer N1? dans le deuxième compartiment (c) et (d), et ainsi de
suite. Avec les deux modèles Pi? décroît d’une manière géométrique.
prédateur était inclinée, mais aussi que le modèle ratio-dépendant est une meilleure
approximation des dynamiques réelles que le modèle proie-dépendant.
1.2.3f
Dérivations mécanistes de la ratio-dépendance
Une critique très fréquente de la théorie ratio-dépendante concerne son introduction
par une justi…cation purement empirique (Murdoch et al. 1998; Abrams 1994). A
l’origine, Arditi & Ginzburg (1989) ont énuméré plusieurs mécanismes potentiels qui
peuvent mener à la ratio-dépendance (échelles de temps différentes entre consommation
et croissance de la population, interférence directe entre prédateurs, existence de refuges
pour la proie, stratégie de recherche non-aléatoire, aspect généraux de l’hétérogénité)
mais sans les développer en détail. Entre temps, quelques-uns de ces mécanismes ont
été analysés de façon mathématique, et on a vu qu’ils peuvent effectivement mener
à la ratio-dépendance dans certains cas. Michalski et al. (1997) et Poggiale et al.
(1998) ont étudié des systèmes avec un refuge pour les proies et où une migration des
proies entre le patch refuge et l’autre patch se produit à une échelle de temps plus
lente que la consommation dans le patch contenant le prédateur (où la prédation se
fait d’une manière proie-dépendante). Ce modèle peut être agrégé mathématiquement
en un modèle proie-prédateur avec un seul patch et une réponse fonctionnelle ratio-
C. Jost
Identi…cation de modèles proie-prédateur (thèse de doctorat)
13f
dépendante. Cosner et al. (in press) a étudié différents types d’organisation spatiale
des prédateurs pendant le fouragement. En fonction de la géométrie de la distribution
des prédateurs (homogène, en patch, sur une ligne, etc.) on obtient plusieurs types de
réponse fonctionnelle, y compris ratio-dépendante.
On peut aussi obtenir cette ratio-dépendance d’une manière mécaniste en revenant
sur la formulation originale proposée par Beddington (1975),
g(N, P ) =
aN
,
1 + ahN + cP 0
(1f.12)
où c est le produit du taux de rencontre entre prédateurs et du temps d’interférence, et
où P 0 = P uP est le nombre de prédateurs total moins une unité (uP ). Beddington a
obtenu cette forme en considérant trois activités possibles pour le prédateur : recherche
de proies, consommation de proies et interférence entre prédateurs (pendant un temps
constant). Malgré quelques problèmes dans la déduction (Ruxton et al. 1992), cette
réponse fonctionnelle est souvent citée dans la littérature, habituellement en écrivant P
au lieu de P 0 . Cependant, si l’on garde P 0 et que l’on réécrit (1f.12) on obtient
g(N, P ) =
(1
aN
.
cuP ) + ahN + cP
Pour c = 0 ceci est la réponse fonctionnelle habituelle de Holling II, tandis que pour
c = 1/uP on obtient la réponse fonctionnelle ratio-dépendante. Le modèle de DeAngelisBeddington est donc un vrai modèle intermédiaire, comme le modèle de Hassell-VarleyHolling (1f.11).
D’autres dérivations mécanistes dans le contexte microbiologique sont mentionnées
dans le chapitre 6 : Fujimoto (1963) a trouvé la ratio-dépendance dans la cinétique des
enzymes, tandis que Characklis (1978) l’a obtenue par un raisonnement fondé sur la
cinétique de saturation appliquée à une croissance limitée par le transfert de masse.
1.2.4f
Des alternatives pour la prédateur-dépendance
Avant de conclure cette section je vais discuter les alternatives possibles à la prédateurdépendance. Revenons au modèle proie-prédateur de base (1f.3). On voit que l’isocline
inclinée du prédateur, telle qu’elle est suggérée par le fait empirique que les abondances à
l’équilibre sont corrélées le long d’un gradient d’enrichissement, pourrait également être
expliquée par une mortalité densité-dépendante du prédateur (P ) (Gatto 1991; Gleeson 1994). L’idée d’une telle densité-dépendance à été avancée à plusieurs reprises dans
des modèles proie-prédateur et dans des modèles plus complexes (DeAngelis et al. 1975;
Steele & Henderson 1981; Steele & Henderson 1992; Edwards & Brindley 1996, voir aussi
tableau A.4). Dans des modèles proie-prédateur, cette mortalité densité-dépendante résulte effectivement en une corrélation entre les abondance à l’équilibre. Par contre, dans
des chaînes trophiques à trois niveaux avec mortalité densité-dépendante dans le niveau
le plus haut, il peut y avoir une corrélation positive ou négative entre les deux premiers
niveaux trophiques (voir 4.B). Des patterns encore plus complexes sont à attendre dans
des chaînes trophiques encore plus longues. La prédateur-dépendance dans la réponse
14f
Identi…cation de modèles proie-prédateur (thèse de doctorat)
C. Jost
fonctionnelle peut donc être considérée comme une explication ’naturelle’ parce qu’elle
prédit cette corrélation positive entre des abondance à l’équilibre le long d’un gradient
d’enrichissement pour tous les niveaux trophiques et indépendamment des valeurs spéci…ques des paramètres. De plus, les données de la réponse fonctionnelle dans Arditi
& Akçakaya (1990) démontrent directement la prédateur-dépendance dans la réponse
fonctionnelle. Je n’ai pas trouvé de données expérimentales en faveur de l’existence
d’une mortalité densité-dépendante.
Une autre alternative est d’abandonner le cadre général (1f.3), en particulier la
conservation de la masse. Un tel modèle proie-prédateur, bien connu, remonte à Leslie
(1948). Ce modèle proie-prédateur conserve la réponse fonctionnelle dans l’équation de
la proie, mais utilise une fonction logistique dans l’équation du prédateur,
dN
N
= r1 (1
) g(N )P
dt
K
dP
P
= (r2 b )N
dt
N
(interaction Lotka-Volterra standard).
g(N ) = aN
(1f.13)
Une extension est d’utiliser une réponse fonctionnelle du type Holling II (1.6). Ce modèle a été analysé par May (1975) et par Tanner (1975), et est connu sous les noms
de modèle de Leslie-May ou de Holling-Tanner. Deux propriétés rendent ce modèle
intéressant: premièrement, l’existence de cycles limite où les populations restent bien
éloignées de l’extinction, et deuxièmement, un lien plus varié entre enrichissement et
stabilité de l’équilibre non-trivial. L’idée d’une densité-dépendance basée sur le rapport entre les abondances de proies et de prédateurs à été généralisée pour des réseaux
trophiques, et est connue sous le nom ‘théorie des réseaux trophiques logistiques’ (Berryman et al. 1995). Dans la littérature, ce type de modèle est aussi souvent quali…é de
ratio-dépendant (Freedman & Mathsen 1993; Hsu & Huang 1995), mais ici cela fait
référence au rapport nombre de prédateurs par proie. Ginzburg (1998) fait valoir que
la conservation de la masse est un élément important dans les réseaux trophiques. Je
ne discuterai dans cette thèse que des modèles qui suivent ce principe.
1.2.5f
Ratio-dépendance : l’état de l’art
Les résultats qui ont été discutés dans les pages précédantes démontrent que la prédateurdépendance est un trait fréquent de la réponse fonctionnelle dans les relations proieprédateur en laboratoire et dans la nature. Or, la verticalité de l’isocline du prédateur,
prédite par des réponses fonctionnelles proie-dépendantes dans le cadre du système
(1f.4), a été contredite plusieurs fois par des données expérimentales. Mais il existe
beaucoup d’alternatives (Hassell & Varley 1969; Hassell & Rogers 1972; Beddington
1975; DeAngelis et al. 1975; Gatto 1991). La ratio-dépendance n’est qu’une possibilité,
mais une possibilité particulièrement simple. En fait, la simplicité est un argument
important car la modélisation comme outil de base en écologie appliquée (gestion des
ressources naturelles, biologie de la conservation, évaluation de l’impact environnemental, etc.) doit être simple à appliquer, avec peu de paramètres à estimer, faute de quoi
les modèles proposés ne vont pas être utilisés par les praticiens. Deux paramètres pour
C. Jost
Identi…cation de modèles proie-prédateur (thèse de doctorat)
15f
la réponse fonctionnelle représentent déjà une limite supérieure, et je ne connais pas
d’étude en écologie appliquée utilisant une réponse fonctionnelle plus complexe. Néanmoins, tandis que les expériences mentionnées (Arditi et al. 1991b; Arditi & Saïah
1992) et la corrélation entre les abondances à l’équilibre le long d’un gradient d’enrichissement démontrent que l’isocline du prédateur n’est souvent pas verticale, et que
l’hétérogénité spatiale diminue la pente de cette isocline, ces deux résultats ne permettent pas de distinguer entre une isocline pentue en général (Hassell & Varley 1969;
DeAngelis et al. 1975) et une isocline passant par l’origine comme elle est prédite par
le modèle ratio-dépendant (voir aussi Figure 1f.1). Arditi & Akçakaya (1990) est le seul
article indiquant que le modèle ratio-dépendant est souvent plus proche de la réalité que
les modèles proie-dépendants (les estimations du paramètre d’interférence m étant majoritairement plus proche de 1 que de 0). Cependant, comme ce paramètre m apparaît
dans un contexte non-linéaire, le fait qu’il soit plus proche de 1 que de 0 n’indique pas
forcément qu’une diagonale est une meilleure approximation de l’isocline du prédateur
qu’une verticale dans le domaine des valeurs d’abondance observées (voir Figure 1f.2 et
4.2).
P
1.2
1
0.8
0.6
0.4
0.2
0.2
0.4
0.6
0.8
1
N
Figure 1f.2: Relation entre la forme de l’isocline du prédateur et le paramètre m
(1f.11). La région ombrée en gris correspond aux valeurs de m estimées dans Arditi
& Akçakaya (1990). Pour de faibles valeurs de m une isocline verticale pourrait
être une approximation plus raisonnable que l’isocline passant par l’origine.
Malgré toutes ces évidences empiriques et théoriques contre les modèles proie-prédateur proie-dépendants, le modèle ratio-dépendant est encore loin d’être accepté comme
cadre théorique valable pour la modélisation de systèmes proie-prédateur ou de systèmes plus complexes. Tandis que certains chercheurs le considèrent comme une pure
absurdité (Abrams 1994; Abrams 1997), d’autres chercheurs ont peur que ce modèle
ne détourne l’attention des formes plus générales de prédateur-dépendance (Murdoch
et al. 1998). D’autres chercheurs encore critiquent le fait que les prédateurs puissent,
théoriquement, persister pour des ressources arbitrairement petites (Yodzis 1994). Mais
ce dernier argument est également valable pour le modèle logistique, et personne n’exige
16f
Identi…cation de modèles proie-prédateur (thèse de doctorat)
C. Jost
l’abandon de la croissance logistique (1f.1) par ex. en faveur de son homologue avec
effet Allee (Allee 1931) (voir section 2.3.1f). De plus, rien n’indique que la disparition
du prédateur lorsque les ressources sont faibles est un résultat déterministe plutôt que
stochastique. La stochasticité démographique ou le ’mutational meltdown’ (accumulation de mutations délétères due à une pression de sélection plus faible par rapport
à la dérive génétique dans des petites populations que dans les grandes, Lynch et al.
1995) peuvent également provoquer une extinction des prédateurs à des niveaux de
ressources faibles. En fait, le raisonnement sous-jacent à la critique de Yodzis a été
développé le plus clairement par Oksanen et al. (1981), et mène à la prédiction que des
environnements plus riches permettent l’existence de chaînes trophiques plus longues.
Cependant, Pimm (1991, p. 224) démontre qu’il n’y a aucune évidence pour une telle
corrélation entre productivité primaire et longeur des chaînes trophiques.
En somme, il reste du travail à faire sur plusieurs points. Premièrement, les modèles
proie-dépendants ou ratio-dépendants doivent être comparés à des données sur un pied
d’égalité, par ex. par des outils de comparaison statistiques (régression non-linéaire dans
un cadre de vraisemblance), ou par la construction d’expériences qui permettent de rejeter les deux types d’isoclines, verticales et diagonales. Deuxièmement, il y a encore des
détails mathématiquement mal connus dans le modèle ratio-dépendant (comportement
dyamique à l’origine, où ce modèle n’est pas dé…ni). Une compréhension mathématique approfondie peut indiquer les faiblesses et les points forts d’un modèle écologique.
Troisièment, d’autres possibilités d’inclure la prédateur-dépendance d’une manière aussi
simple que dans les modèles ratio-dépendants doivent être étudiées et confrontées avec
l’approche ratio-dépendante. Le paramètre d’interférence m proposé par Hassell & Varley (1969) est un tel candidat, la réponse fonctionnelle proposée par Ashby (1976) (voir
Tableau A.3) en est un autre. Dans cette thèse, je vais contribuer principalement au
premier et deuxième points. D’autres possibilités d’inclure la prédateur-dépendance ne
seront pas étudiées car elles demandent plus de paramètres pour avoir la même gamme
de comportements dynamiques que le modèle ratio-dépendant (voir section 2.3.1f et
chapitre 8).
1.2.6f
Une petite digression : à quoi se refère le terme de
“densité” dans un réponse fonctionnelle densité-dépendante?
Il y a dans la littérature une certaine confusion concernant l’utilisation du terme ‘densitédépendance’ dans le contexte de modèles proie-prédateur, car le terme de densité peut
faire référence à l’abondance du prédateur ou à celle de la proie. Une approche raisonnable
a été choisie par Ruxton et collaborateurs, qui quali…ent une réponse fonctionnelle
prédateur-dépendante de réponse fonctionnelle densité-dépendante (Ruxton & Gurney
1992; Ruxton 1995) en raison du lien intrinsèque entre la réponse fonctionnelle et le
prédateur (pas de réponse fonctionnelle sans prédateur), et aussi parce que la densitédépendance désigne, dans son sens original, un effet délétère sur la croissance d’un
organisme. Cependant, d’autres auteurs se réfèrent à la densité de la proie en utilisant
le terme ’density-dependent predation’ (Sih 1984; Trexler 1988; Moore 1988; Mansour
& Lipcius 1991; Possingham et al. 1994). Cette utilisation a ses origines dans la publi-
C. Jost
Identi…cation de modèles proie-prédateur (thèse de doctorat)
17f
cation qui a dé…ni les termes de ‘réponse fonctionnelle’ et ‘réponse numérique’ pour la
première fois, Solomon (1949, p.16): “To be density-dependent, the enemy must take a
greater proportion of the population as the host density increases”. Solomon lui-même
s’appuie sur Varley (1947) et Nicholson (1933) pour cette utilisation. En dépit de cette
antériorité historique, je considère ceci comme une utilisation fausse du terme, de nature à augmenter la confusion. Pour éviter toute confusion, on peut appeler une réponse
fonctionnelle qui ne dépend que de la proie une réponse fonctionnelle proie-dépendante.
Quand cette fonction dépend aussi de la densité du prédateur, on peut parler d’une
réponse fonctionnelle prédateur-dépendante.
18f
Identi…cation de modèles proie-prédateur (thèse de doctorat)
C. Jost
Chapitre 2f
Un cas de sélection de modèle
Comparer les modèles pouvant être inclus dans la classe des modèles proie-dépendants et
ratio-dépendants en général est une tâche difficile, peut-être même impossible. Une des
approches possibles est de choisir un représentant paramétrique de chaque classe, tous
deux de même complexité (càd ayant le même nombre de paramètres et la même gamme
de comportements dynamiques), et de les comparer à des données écologiques d’une
manière qualitative (analyse de stabilité, réaction à des changements de paramètres,
dynamique temporelle) et quantitative (régression non-linéaire, vraisemblance). Une
meilleure concordance entre l’un des modèles et les données peut être interprétée comme
un signe que la prédateur-dépendance dans la réponse fonctionnelle est plus ou moins
importante. Les données disponibles sont des dynamiques temporelles qualitatives d’un
système proie-prédateur (dynamique du plancton dans les lacs d’eau douce, section
2.2f) ou séries chronologiques de systèmes proie-prédateur simples (protozoaires élevés
en laboratoire), de complexité moyenne (arthropodes élevés en laboratoire), ou plus
complexes (plancton dans les eaux douces), voir section 2.5f).
2.1f
Les modèles proie-prédateur potentiels
Je vais analyser deux modèles proie–prédateur. Tous deux ont une croissance logistique
de la proie (en l’absence de prédateur) et un taux de mortalité du prédateur constant. La
réponse fonctionnelle est du type Holling II (1.6), tandis que le modèle ratio-dépendant
est aussi de cette forme, mais avec le ratio N/P (proie par prédateur) comme argument,
dN
dt
dP
dt
N
)N
K
=
r(1
=
eg (N, P )P
g(N, P )P
P
(2f.1)
aN
1 + ahN
g(N, P ) →
N/P
.
1 + hN/P
Les deux modèles ont six paramètres. On verra dans le chapitre 4 qu’ils ont trois équilibres possibles et que les dynamiques globales sont la coexistence stable, la coexistence
19f
20f
Identi…cation de modèles proie-prédateur (thèse de doctorat)
C. Jost
instable (cycle limite), l’extinction du prédateur seul ou l’extinction des deux populations (ce dernier cas n’existe qu’approximativement dans le modèle proie-dépendant,
i.e., une oscillation d’un cycle limite de grande amplitude amenant les dynamiques près
de l’extinction). Ces dynamiques se trouvent aussi dans les séries chronologiques avec
lesquelles je vais comparer les modèles.
La densité-dépendance dans la croissance de la proie est nécessaire pour deux raisons:
(1) que la population de la proie n’explose pas en absence d’un prédateur, et (2) parce
qu’une croissance malthusienne combinée avec une réponse fonctionnelle bornée résulterait toujours en un équilibre non-trivial instable. Il existe des fonctions de croissance
alternatives (Gompertz 1825; Rosenzweig 1971), et j’ai choisi la croissance logistique
simplement parce qu’elle est largement acceptée. Les deux réponses fonctionnelles sont
bornées (saturation de la consommation), un trait dont le réalisme écologique est incontesté. Il existe des alternatives proie-dépendantes à la réponse fonctionnelle du type
Holling II qui ont aussi deux paramètres (Ivlev 1961; Watt 1959), et j’ai choisi la forme
Holling type II parce qu’elle est la réponse fonctionnelle la plus répandue et la mieux
connue. Il existe aussi des alternatives prédateur-dépendantes à la réponse fonctionnelle
ratio-dépendante, telles que la réponse fonctionnelle linéaire du type Hassell-Varley,
g(N, P ) = P
m
N,
ou la forme proposée par Ashby (1976) ou Hassell & Rogers (1972),
g(N, P ) =
1
aN
.
1 + ahN P
Tandis que ces deux alternatives n’ont que deux paramètres, comme le modèle ratiodépendant, elles ont toutes les deux le désavantage de ne pas être bornées pour P → 0,
et que dans le cadre (2f.1) elles ne présentent pas d’équilibre non-trivial instable ou
des cycles limites stables. Nous pouvons argumenter que le modèle ratio-dépendant
ressemblant au modèlle Holling II représente la plus simple des réponses fonctionnelles
ayant des dynamiques similaires à celles avec une réponse fonctionnelle proie-dépendante
du type Holling II.
2.1.1f
Petite remarque : modèle discret et modèle continu
Les deux modèles développés dans la section précédente sont tous deux des modèles en
temps continu (équations différentielles). Ces modèles s’appliquent à des grandes population avec des générations chevauchantes. L’autre extrême serait le cas de populations
consistant en une seule génération sans chevauchement entre les générations (par ex.
des plantes annuelles). Ces systèmes devraient évoluer à pas de temps discrets,
Nt+1 = F (Nt , Pt )
Pt+1 = G(Nt , Pt )
(pour plus d’informations sur les modèles de ce type voir May (1976a) ou Begon et al.
(1996a)). L’approche ratio-dépendante a rarement été utilisée dans un contexte discret,
C. Jost
Identi…cation de modèles proie-prédateur (thèse de doctorat)
21f
les seuls exemples à ma connaissance en sont Carpenter et al. (1993) et Carpenter et al.
(1994) (utilisation de modèle discret pour des raisons techniques dans leur analyse de
séries chronologiques). Les données réelles utilisées dans cette thèse peuvent être caractérisées par des générations chevauchantes avec des grandes populations, elles concordent donc mieux avec des modèles en temps continu. C’est la raison pour laquelle
j’ai utilisé et comparé seulement des équations différentielles, ce qui est numériquement plus lourd, mais plus facile d’un point de vue théorique. L’utilité de l’approche
ratio-dépendante pour des modèles discrets n’a pas été explorée a l’heure actuelle.
2.2f
Comparaison qualitative des modèles
Cette section est un sommaire des idées et des résultats de l’analyse présentée au chapitre
4.
Les dynamiques temporelles les mieux connues en dehors de systèmes de laboratoire proviennent probablement du plancton des lacs d’eau douce en zone tempérée. En
étudiant les dynamiques détaillées du phyto- et du zooplancton dans 24 lacs, Sommer
et al. (1986) ont dévéloppé un modèle verbal de ces dynamiques en 24 énoncés, nommé
modèle PEG (PEG vient de ‘Plankton Ecology Group’). Ces dynamiques peuvent être
représentées sommairement comme dans la Figure 4.1. Au printemps, les populations
de phyto- et de zooplancton commencent à croître depuis les niveaux bas de l’hiver. Le
phytoplancton arrive à un sommet printanier très élevé, suivi par un fort déclin vers des
niveaux très bas dûs à une surconsommation par le zooplancton herbivore et à un épuisement des nutriments. Ce déclin est appelé la phase des eaux claires (CWP, de ‘Clear
Water Phase’) parce que l’eau apparaît très claire en raison des abondances très basses
du phytoplancton. Après ce déclin, le phytoplancton augmente de nouveau jusqu’à des
niveaux comparables à ceux du maximum printanier, dominé largement par des espèces
non-comestibles (…lamenteuses, toxiques, avec des structures de défense), et il y reste
durant l’été jusqu’à ce que la température et la lumière décroissantes engendrent un
nouveau déclin de la population en automne. Le zooplancton suit ces dynamiques avec
un certain retard et avec des maxima et des minima moins distincts. Ces dynamiques
semblent plus au moins indépendantes de l’état trophique du lac. Seules les amplitudes
augmentent et le maximum printanier est atteint plus tôt en conditions eutrophes.
2.2.1f
Interprétation dans le context proie-prédateur
Les dynamiques entre le printemps et l’automne sont interprétées dans le modèle PEG
comme résultant d’interactions trophiques. Dans le cas le plus simple, le phytoplancton
est considéré comme étant la proie et le zooplancton (majoritairement Daphnia sp.) le
prédateur. Les dynamiques temporelles correspondent à des dynamiques transitoires
qui commencent avec des conditions initiales basses et convergent vers un équilibre
durant l’été après quelques oscillations fortes. Le chapitre 4 étudie si les deux modèles
considérés dans cette thèse (2f.1) peuvent expliquer ces dynamiques, et s’il existe des
différences entre leur pouvoir prédictif.
22f
Identi…cation de modèles proie-prédateur (thèse de doctorat)
C. Jost
Dans une première étape, nous analysons les dynamiques globales des deux systèmes
proie-prédateur. Une première observation concerne l’équilibre de la proie et la façon
dont il change avec un enrichissement (eutrophisation, modélisée avec la capacité de
charge K ): cet équilibre est indépendant de la capacité de charge dans le modèle proiedépendant, tandis qu’il est proportionnel à la capacité de charge dans le modèle ratiodépendant. Les observations dans des lacs naturels révèlent une forte corrélation entre le
contenu en nutriments et l’abondance moyenne du phytoplancton (McCauley et al. 1988;
Mazumder & Lean 1994), et sont donc en accord avec les prédictions du modèle ratiodépendant. Pour étudier les dynamiques transitoires, nous construisons tout d’abord
les isoclines. Les deux modèles ont une isocline de la proie bossue et une isocline
du prédateur droite, la différence étant que le modèle proie-dépendant a une isocline
du prédateur verticale et le modèle ratio-dépendant une isocline inclinée et passant par
l’origine (Figure 4.2). Il y a au maximum trois équilibres : l’origine (0, 0), le système sans
prédateur (K, 0) et un équilibre non-trivial. Ici, on se concentrera sur ce dernier équilibre
(non-trivial) qui devrait être localement stable et atteint avec des oscillations pour
concorder avec le modèle PEG. Après une transformation en système sans dimensions,
on peut identi…er dans l’espace des paramètres (simpli…é) la région qui permet ces
dynamiques, voir Figure 4.3. La différence principale entre les deux modèles est encore
la réaction à un enrichissement (augmentation de K ) : déstabilisation dans le modèle
proie-dépendant, simple augmentation des amplitudes dans le modèle ratio-dépendant.
Ceci rend le modèle ratio-dépendant plus proche des dynamiques observées, qui n’ont
révélé aucun signe de déstabilisation dans des lacs eutrophes. Toutefois, on a aussi
observé que le maximum printanier est atteint plus tôt dans des systèmes eutrophes.
Le modèle ratio-dépendant ne peut expliquer cette observation que si l’eutrophisation
entraîne aussi une augmentation du taux de croissance r.
En dépit du faible avantage du modèle ratio-dépendant, aucun des deux modèles
ne peut reproduire les dynamiques du modèle PEG : soit il y a une forte première
oscillation (créant la CWP), mais dans ce cas le système continue ses oscillation durant
l’été, soit le système se stabilise au début de l’éte, mais dans ce cas il n’y a pas de CWP
(Figure 4.5). Apparemment, les deux systèmes proie-prédateurs sont trop simples pour
expliquer ces dynamiques.
McCauley et al. (1988) suggèrent que plusieurs paramètres changent durant la saison:
le taux de croissance r augmente en raison d’une température de l’eau plus élevée,
le taux d’attaque a (ou ) décroît en raison d’une augmentation de la proportion
d’algues non-comestibles,
la mortalité du prédateur croît en raison d’une prédation d’un niveau trophique
supérieure (sur les alevins) en été.
Ces changements des paramètres suffisent pour générer les dynamiques désirées avec
les deux modèles (voir Figure 4.6 pour un exemple).
Nous pouvons conclure que les deux modèles sont trop simples pour reproduire les
dynamiques du modèle PEG si les paramètres restent constants, mais que les deux mod-
C. Jost
Identi…cation de modèles proie-prédateur (thèse de doctorat)
23f
èles peuvent les reproduire si un ou plusieurs des paramètres peuvent changer durant la
saison. Le seul trait qui distingue les deux modèles est leur réaction à un enrichissement
(si on traduit cet enrichissement uniquement par une augmentation de la capacité de
charge K ). Dans ce cas, les deux modèles ont des traits qui sont incompatibles avec
le modèle PEG: le modèle proie-dépendant ne peut pas expliquer l’augmentation du
phytoplancton, et le modèle ratio-dépendant ne peut pas expliquer que le maximum
printanier soit atteint plus tôt (ou un changement dans les périodes des oscillations en
général). Néanmoins, les deux modèles concordent avec le modèle PEG si l’enrichissement augmente aussi le taux de croissance r.
2.2.2f
Ajout d’un niveau trophique
Les résultats discutés dans cette section ne font pas partie de l’article dans le chapitre 4.
Ils forment une évaluation préliminaire de la façon dont le travail du chapitre 4 pourrait
être continué.
Les modi…cations saisonnières des paramètres mentionnées au paragraphe précédent
sont essentiellement dues à un changement de l’in‡uence d’un autre niveau trophique sur
le système phytoplancton-zooplancton (un prédateur supérieur consommant Daphnia,
d’où une augmentation de la proportion d’algues non comestibles). Au lieu de changer
la valeur d’un paramètre, il est également possible d’inclure explicitement ce niveau
trophique supplémentaire dans le modèle. L’ajout d’un prédateur supérieur C aboutit
au système
dN
dt
dP
dt
dC
dt
= f (N )N
g1 (N, P )P
= e1 g1 (N, P )P
g2 (P, C )C
= e2 g2 (P, C )C
C.
(2f.2)
Bernard & Gouzé (1995) ont développé une méthode permettant d’étudier la succession possible d’extrema pour ce type de modèles de chaîne trophique. Si l’on suit la
dynamique du modèle PEG et que l’on marque chaque niveau trophique avec un + s’il
augmente et avec un s’il diminue, on peut représenter la dynamique du modèle PEG
par la succession suivante de vecteurs (voir l’Annexe D pour plus de détails),
’
+
+
- +
- +
- +
- +
$
sign(N˙ )
- + := sign(P˙ )
sign(C˙ )
&
%
Grâce à la méthode de Bernard & Gouzé (1995), nous pouvons tester si des chaînes
trophiques proie- et ratio-dépendantes présentent bien cette succession d’événements.
24f
Identi…cation de modèles proie-prédateur (thèse de doctorat)
C. Jost
Cette analyse montre que les deux types de modèles de chaîne trophique présentent
exactement la même succession d’extrema et peuvent tous deux prédire la succession
décrite ci-dessus (voir l’Annexe D pour plus de détails).
Il est possible de faire une analyse plus quantitative en paramétrisant directement le
système (2f.2) avec une croissance logistique de la proie et une réponse fonctionnelle soit
proie- soit ratio-dépendante (de la forme 2f.1) et de montrer qu’il existe des valeurs de
paramètres qui produisent la dynamique désirée. Ceci constitue une forme de validation
de modèle dans la mesure où l’on véri…e que le modèle proposé peut effectivement
reproduire une dynamique observée par ailleurs. Nous supposons que l’abondance du
prédateur supérieur de cette chaîne augmente au cours de la saison. La Figure 2f.1
montre un exemple de dynamique produite par les deux types de réponse fonctionnelle.
Dans une telle chaîne trophique à trois niveaux, le modèle ratio-dépendant conserve
son avantage par rapport au modèle proie-dépendant, en particulier la corrélation entre
les abondances d’équilibre des proies et des prédateurs sur un gradient de richesse du
système.
Figure 2f.1: Exemples numériques d’une chaîne trophique à trois niveaux, ‘proieprédateur-top-prédateur’, avec une réponse fonctionnelle proie-dépendante (a) et
ratio-dépendante (b). Les éléments caractéristiques du modèle PEG sont bien
démontrés. ———— : dynamique de la proie; – – – : dynamique du prédateur;
— — — : dynamique du top-prédateur.
Il est également possible de modéliser explicitement les deux types de proies, comestibles (Ne ) et non comestibles (Ni ), qui sont en compétition les unes avec les autres,
dNe
Ne
= re (1
)Ne ce Ne Ni g(Ne , P )P
dt
Ke
dNi
Ni
= ri (1
)Ni ci Ni Ne
(2f.3)
dt
Ki
dP
= eg (Ne , P )P P,
dt
la réponse fonctionnelle étant à nouveau du type (2f.1). Un scénario raisonnable pour
un tel système est que Ne est un compétiteur supérieur par rapport à Ni (Agrawal
1998) mais ne parvient pas à l’exclure en raison de la prédation qu’il subit par ailleurs.
La Figure 2f.2 montre une réalisation numérique pour un tel système. De nouveau, la
dynamique du modèle PEG est clairement reproduite. Dans ce scénario, les deux types
de réponses fonctionnelles prédisent une corrélation entre les abondances de proies et
les abondances de prédateurs à l’équilibre.
C. Jost
Identi…cation de modèles proie-prédateur (thèse de doctorat)
(a)
time
25f
(b)
time
Figure 2f.2: Exemple numérique d’un système comportant une proie comestible,
une proie non-comestible et un prédateur. Le modèle proie-dépendant (a) et le
modèle ratio-dépendant (b) présentent les charactéristiques du modèle PEG. —
——— : dynamique de la proie comestible; – – – : dynamique de la proie noncomestible; — — — : dynamique du prédateur.
2.2.3f
Conclusions
En somme, ni les modèles proie-prédateur à réponse fonctionnelle proie-dépendante
ni les modèles ratio-dépendants ne comprenant que du phytoplancton et zooplancton
ne sont capables de prédire la dynamique saisonnière de ces deux niveaux trophiques,
sauf si l’on autorise des variations saisonnières des valeurs des paramètres (auquel cas
les deux types de modèles peuvent prédire cette dynamique). Toutefois, le modèle
ratio-dépendant semble plus réaliste du point de vue des propriétés suivantes : 1. les
biomasses d’équilibre du phytoplancton et du zooplancton augmentent lorsque la productivité primaire augmente, 2. la productivité n’affecte pas l’équilibre du système.
Pour obtenir le pattern attendu, avec une phase des eaux claires (CWP) bien distincte
et un équilibre stable rapidement atteint en été, les modi…cations suivantes peuvent être
apportées au modèle ratio-dépendant à deux niveaux trophiques : 1. variations saisonnières des paramètres du phyto- et du zooplancton, 2. introduction d’un troisième
niveau trophique (zooplancton carnivore ou poissons), 3. introduction d’un facteur de
mortalité du phytoplancton autre que la prédation (sédimentation).
2.3f
Dynamiques de bord : quelle importance ?
Cette section résume les idées et les résultats de l’analyse développée au chapitre 5.
Dans la section précédente nous avons comparé les deux systèmes proie-prédateur
(2f.1) à des données aboutissant à un équilibre stable non trivial. Le modèle ratiodépendant permet également d’aboutir à une extinction déterministe (Arditi & Berryman 1991). Cependant, le modèle ratio-dépendant n’est pas directement dé…ni à l’origine car cela implique une division par zéro. Toutefois, si l’on suppose une réponse
fonctionnelle bornée (comme dans les systèmes 2f.1) alors l’origine est effectivement un
point d’équilibre, simplement on ne peut pas appliquer à ce point une analyse d’équilibre standard, ce qui explique que, pendant longtemps, la dynamique autour de ce point
d’équilibre est restée mal comprise. Le fait que cette dynamique soit mal comprise a
26f
Identi…cation de modèles proie-prédateur (thèse de doctorat)
C. Jost
été utilisé de façon répétée pour critiquer le concept de ratio-dépendance (Yodzis 1994;
Abrams 1997), et des résultats mathématiques faux ou imprécis ont été publiés (Getz
1984; Freedman & Mathsen 1993).
En collaboration avec Ovide Arino j’ai analysé en détail le comportement analytique
à proximité de l’origine. La façon la plus parlante d’exposer les résultats est de les
représenter graphiquement. La Figure 5.1 montre des portraits de phase possibles du
modèle ratio-dépendant (2f.1). Si les cas 5.1a et 5.1c ne présentent pas de problèmes
mathématiques particuliers, la situation représentée en 5.1b est mal comprise. En transformant les variables d’état (en étudiant les systèmes (N/P, P ) et (N, P/N ) au lieu du
système (N, P )) on peut montrer que, pour certaines valeurs des paramètres l’origine se
comporte en effet comme un point de selle (Figure 5.2). Toutefois, lorsque l’efficacité
des prédateurs augmente (ce qui est matérialisé par le fait que la pente de l’isocline
augmente) l’origine devient attractive pour les trajectoires où ‘N va plus vite vers 0
que P ’. Ceci se produit déjà pour des valeurs de paramètres pour lesquelles l’équilibre
non trivial est encore localement stable (Figure 5.3), ce qui divise l’espace de phases en
deux bassins d’attraction. L’existence de telles trajectoires est prouvée pour les modèles
ratio-dépendants dans le cas général, à la seule condition que la réponse fonctionnelle
soit croissante et bornée et que la fonction de croissance ‘per capita’ de la proie soit
une fonction décroissante de l’abondance des proies. Une augmentation supplémentaire de l’efficacité du prédateur peut conduire à une déstabilisation de l’équilibre non
trivial, ce qui aboutit à un cycle limite stable (Figure 5.4). Toutefois, l’existence de
ce cyle limite n’est pas encore prouvée du point de vue analytique. Le problème est
que l’origine devient attractive avant que l’équilibre non trivial ne devienne instable.
Aussi la construction d’un ensemble invariant, nécessaire à l’application du théorème de
Poincaré-Bendixson pour prouver l’existence d’un cycle limite stable, nécessiterait-elle
une connaissance précise de la séparatrice entre les deux bassins d’attraction. Lorsque
l’efficacité du prédateur augmente plus encore, il se produit une bifurcation homoclinique (détectée de façon numérique) après laquelle l’origine devient attractive pour toutes
les conditions initiales sauf l’équilibre (instable) non trivial lui-même (Figure 5.5).
Gragnani (1997) attire l’attention sur le fait que cette extinction déterministe n’est
pas un comportement générique, en arguant que, si l’on considère un modèle intermédiaire du type DeAngelis-Beddington (1f.12), alors l’origine reste un point de selle
pour toute valeur des paramètres sauf celles qui rendent ce modèle ratio-dépendant.
L’auteur ne précise pas que cette argumentation repose entièrement sur le choix (arbitraire) du modèle intermédiaire (d’ailleurs, la bifurcation homoclinique, autrement
dit l’attractivité globale de l’origine alors même qu’il existe un équilibre non trivial, a
aussi entièrement échappé à l’auteur). Si l’on choisit la réponse fonctionnelle de HassellVarley-Holling (1f.11) comme modèle intermédiaire, le tableau s’en trouve quelque peu
modi…é. Callois (1997) a étudié ce modèle en détail et a montré que l’extinction déterministe se produit pour m 1, tandis que l’origine est un point de selle pour m < 1. Il
semble donc qu’en effet l’extinction n’est pas générique pour m 1.
C. Jost
2.3.1f
Identi…cation de modèles proie-prédateur (thèse de doctorat)
27f
Quelques autres modèles offrant l’extinction
Les deux sections suivantes ne seront pas détaillées au chapitre 5. Elles présentent des
idées supplémentaires sur l’extinction déterministe et le contrôle biologique.
L’extinction conjointe de la proie et du prédateur est habituellement interprétée comme
étant le résultat d’événements stochastiques. Cependant, le fait que cette extinction se
produise de façon répétée dans les systèmes prédateur-proie de laboratoire (Gause 1934;
Luckinbill 1973; Veilleux 1979) suggère qu’il pourrait également s’agir d’un phénomène
déterministe. En dehors du modèle ratio-dépendant analysé plus haut, je n’ai trouvé
que deux autres systèmes proie-prédateur simples qui permettent ce type d’extinction
déterministe. Le premier est un modèle avec un fort effet Allee dans la dynamique de la
proie (le taux de croissance de la population devenant négatif en dessous d’une valeur
seuil de l’effectif , Allee 1931) et une réponse fonctionnelle de type Lotka-Volterra (de
Hastings 1997)
dN
dt
dN
dt
= r(1
= eaN P
N
N
)N (
K
1)
K
K
aNP
P.
Notons que, dans le cas général, l’existence d’un effet Allee signi…e simplement que le
taux de croissance par individu est maximum non pas à l’origine mais pour une certaine
valeur positive de l’effectif des proies (Edelstein-Keshet 1988). Le fort effet Allee auquel
je fais référence signi…e que le taux de croissance doit même devenir négatif aux faibles
densités (voir Figure 2f.3). Seul un effet Allee fort permet l’extinction de la proie. Le
second modèle inclut une croissance logistique de la proie et une réponse fonctionnelle
prédateur-dépendante, telle que celles proposées par Hassell & Rogers (1972) ou Ashby
(1976),
dN
dt
dP
dt
aN
1
N
P
)N
K
1 + ath N P
aN
1
P P.
= e
1 + ath N P
= r(1
Ce modèle est entièrement contrôlé par le donneur et peut avoir deux équilibres non
triviaux (dont l’un est stable), auquel cas l’origine devient attractive pour n’importe
quelles conditions initiales pour peu que l’abondance des proies soit suffisamment faible.
2.3.2f
Y-a-t-il un paradoxe du contrôle biologique ?
L’analyse présentée au chapitre 5 a révélée une autre caractéristique intéressante des
modèles ratio-dépendants : le fait qu’une coexistence stable entre proies et prédateurs
est possible pour des valeurs d’équibre de l’effectif des proies aussi proches de zéro que
l’on veut. Ceci les distingue des modèles proie-dépendants où des valeurs basses de
l’abondance d’équilibre des proies soit conduisent à une déstabilisation du système, soit
exigent une abondance très élevée des prédateurs (Arditi & Berryman 1991). En réponse
28f
Identi…cation de modèles proie-prédateur (thèse de doctorat)
C. Jost
Figure 2f.3: Quelques cas typiques du taux de variation de la croissance de la proie
f (N ): (a) croissance logistique, (b) effet Allee faible et (c) effet Allee fort. K est
la capacité de charge, N la densité de la proie.
à Arditi & Berryman (1991), Åström (1997) a indiqué qu’une densité-dépendance nonlinéaire dans la fonction de croissance de la proie, de la forme d’une croissance logistique (introduite en écologie par Gilpin & Ayala 1973, mais originellement employée
par Richards 1959),
!
N
f (N ) = r 1
,
K
pouvait stabiliser les systèmes proie-prédateur pour des faibles effectifs d’équilibre des
proies. Toutefois, on ne dispose que de peu d’estimations du paramètre . Åström
(1997) cite des études préliminaires qui semblent indiquer que cette valeur se situe
en-dessous de 1 lorsque les proies sont des insectes, et au-dessus de 1 lorsqu’il s’agit de
mammifères. L’effet stabilisant sur l’équilibre des proies décrit par Åström ne se produit
que pour < 1. Toutefois, une faible valeur de réduit aussi le taux de croissance des
populations lorsque les effectifs sont faibles, ce qui ne correspond pas vraiment à un
ravageur typique.
En fait, le problème du contrôle biologique a été abordé précédemment par Beddington et al. (1978). Ceux-ci ont dé…ni un paramètre q, qui représente le rapport
entre l’abondance d’équilibre (stable) de la proie en coexistence avec le prédateur divisé
par l’abondance d’équilibre de la proie en l’absence du prédateur. Une faible valeur de
q (< 0.01) indique donc un contrôle biologique efficace de la proie par son prédateur.
Beddington et al. (1978) ont trouvé des valeurs de q comprises entre 0.002 et 0.03 sur
le terrain, et entre 0.1 et 0.5 en laboratoire. Ils ont exploré différentes façons possibles
dont les modèles proie-prédateur ou hôte-parasite pourraient être adaptés pour être en
accord avec ces observations :
le fait de donner une forme sigmoïde à la réponse fonctionnelle n’a pas d’effet sur
q,
le fait de tenir compte de l’interférence mutuelle à la manière de DeAngelisBeddington (1f.8) donne des valeurs de q de 0.1 au minimum,
le fait d’incorporer une aggrégation spatiale des parasitoïdes avec une distribution
hétérogène des hôtes rend possible des valeurs de q inférieures à 0.1.
C. Jost
Identi…cation de modèles proie-prédateur (thèse de doctorat)
29f
Dans un article précédent, Free et al. (1977) avaient montré que l’aggrégation spatiale, observée d’un point de vue global, conduit à des taux d’attaque a réduits (comme
l’interférence entre prédateurs), un phénomène que les auteurs ont nommé ‘pseudo-interference’. C’est la raison pour laquelle Beddington et al. (1978) ont modélisé le dernier
élément de la liste ci-dessus avec le modèle d’interférence de Hassell-Varley (1f.7) fondé
sur des estimations du paramètre d’interférence m comprises entre 0.1 et 0.8, et ont
obtenu des valeurs de q jusqu’à 0.004. En utilisant la réponse fonctionnelle de HassellVarley-Holling (1f.11), Arditi & Akçakaya (1990) avaient montré que le paramètre m
pouvait même être plus élevé, permettant ainsi d’obtenir des valeurs de q plus faibles.
Ainsi que nous l’avons déjà indiqué, le modèle de Hassell-Varley-Holling devient ratiodépendant pour m = 1, ce qui suggère que ce modèle est une description très grossière
des problèmes de contrôle biologique.
Identi…ons la région de l’espace des paramètres du modèle ratio-dépendant qui permet de telles valeurs basses de l’abondance des proies à l’équilibre. Pour simpli…er cette
analyse, nous utiliserons la forme sans dimensions du modèle ratio-dépendant introduite
au chapitre 4,
dN
dt
dN
dt
= R(1
=
N
)N
S
SN
P
P + SN
SN
P
P + SN
QP,
avec R = rh/e, S = h/e et Q = h/e. S est la valeur d’équilibre de la proie en
l’absence de prédateurs. Nous cherchons donc à identi…er toutes les combinaisons de
paramètres qui donnent un équilibre des proies N ? bas et stable,
N ? < S.
Un raisonnement algébrique permet de montrer que ceci est possible pour tout S remplissant la condition suivante
(1
)
R
1
Q
< S < min
R
1
Q
,
Q
Q2 + R
1 Q2
La Figure 2f.4 illustre la région de l’espace des paramètres qui remplit cette condition.
Approximativement, il faut donc que S 1 RQ , ce qui signi…e, pour les paramètres
d’origine, que
er
.
e h
Cette condition peut être examinée sur des données réelles. Bernstein (1985) a étudié le
système proie-prédateur Tetranychus urticae (proie) avec Phytoseiulus persimilis (prédateur) et ont paramétrisé un système proie-prédateur où chaque population est structurée en trois stades. Il est possible d’utiliser ses estimations pour paramétriser le
modèle ratio-dépendant (la manière dont le paramètre est désigné dans la publication
originale est indiquée entre parenthèses ) : r = 5.34d 1 (r1 ), = 0.0433d 1 (2 ), e =
30f
Identi…cation de modèles proie-prédateur (thèse de doctorat)
0.159 (min {rm , r2 }) (oeufs pondus par oeufs consommés) et h 0.1d
Ces valeurs donnent
cm2
m2
= 0.00055 .
5.49
d
d
1
C. Jost
(Ki /j or bi /i ).
Cette valeur est proche de celle de = 0.0002m2 /d rapportée par (Hassell 1978b) pour
ce système proie-prédateur. Bien entendu, ceci n’est qu’un résultat préliminaire, mais il
est encourageant de voir que le modèle ratio-dépendant peut se révéler utile en matière
de contrôle biologique.
(1- )R/(1-Q)
Figure 2f.4: La région hâchurée dans l’espace des paramètres R S (pour un Q < 1
…xé) indique les valeurs des paramètres pour lesquelles le prédateur abaisse l’effectif
de la proie jusqu’à un équilibre stable, une fraction au-dessous du niveau que la
proie atteindrait en l’absence du prédateur. Les hâchures verticales indiquent la
région où l’origine se comporte comme un point de selle, tandis que, dans la région
avec les hâchures horizontales, l’origine est attractive pour certaines conditions
initiales.
Il se pourrait également que les systèmes proie-prédateur ne présentent pas d’équilibre stable dans le cadre du contrôle biologique, mais plutôt une dynamique de métapopulation avec de fréquentes extinctions locales. Luck (1990) rapporte (sur la base
de données extraites de Murdoch et al. 1985) que “sur 9 projets de contrôle biologique
couronnés de succès, 1 seul présente une N ? stable, tandis que les autres peuvent s’expliquer par une extinction locale”. Là aussi, de telles extinctions locales sont possibles
avec une description des dynamiques locales à l’aide d’un modèle ratio-dépendant. La
condition sur les paramètres est semblable à celle trouvée plus haut,
S
R
1
Q
or S >
R
1
Q
,
et les types d’extinction sont illustrés dans les Figures 5.5 et 5.1c.
C. Jost
2.4f
31f
Identi…cation de modèles proie-prédateur (thèse de doctorat)
Les microbiologistes l’ont fait
Cette section résume les idées et les résultats de l’analyse développé au chapitre 6.
En fait, alors que Arditi & Ginzburg (1989) ont été les premiers à proposer le modèle
ratio-dépendant en tant que concept général (avec toutes ses implications du point
de vue des régulations par le bas ou par le haut), Getz (1984) a publié un modèle
ratio-dépendant spéci…que plusieurs années avant eux. L’idée elle-même, selon laquelle
la consommation (et avec elle la croissance d’un organisme) est fonction du rapport
entre l’abondance des consommateurs et celle de leurs ressources se trouve dans la
littérature microbiologique dès Cutler & Crump (1924). Il est intéressant de voir qu’une
fonction ratio-dépendante spéci…que de la forme (1f.10) a été proposée en microbiologie
comme fonction de croissance (un concept semblable à celui de réponse fonctionnelle)
par Contois (1959). Depuis, cette fonction a servi comme alternative à l’équivalent de la
fonction de Holling type II en microbiologie, la célèbre fonction de croissance de Monod
(1942). A l’aide du Science Citation Index, j’ai consulté l’ensemble des articles qui
citaient le travail de Contois durant les 40 dernières années a…n de détecter les résultats
qui pouvaient être intéressants dans le contexte écologique.
La majorité du travail expérimental a été effectuée dans des chémostats qui sont
généralement décrits de la façon suivante
dN
dt
dP
dt
= D(N0
N)
= (N, P )P
Y (N, P )P
DP
(2f.4)
avec un substrat N , une concentration du substrat entrant dans le système N0 , un consommateur P , un taux de dilution D et un rendement Y . En comparant cette équation à
notre système proie-prédateur standard (1f.4), on voit qu’il existe une différence subtile
entre les deux façons dont la consommation est modélisée : les microbiologistes partent
d’une fonction de croissance (appelée réponse numérique en écologie) et considèrent que
la fonction de consommation du substrat (notre réponse fonctionnelle) est proportionnelle à la constante de rendemant Y , alors que les écologistes partent de la réponse
fonctionnelle et considèrent que la réponse numérique est proportionnelle à l’efficacité
de conversion e. Il y a donc une différence d’échelle entre les deux concepts,
g(N, P ) = Y (N, P ).
Cette différence est toutefois sans importance pour la discussion ci-après.
Le taux de croissance (N, P ) est habituellement modélisé grâce à l’équation proposée par Monod (1942),
(N, P ) = max
N
Ks + N
avec un taux de croissance maximum max et une constante de demi-saturation Ks (notons qu’avec a = /K et h = 1/ on obtient la réponse fonctionnelle de Holling type
II classique (1f.6)). Avec une telle fonction de croissance, le système (2f.4) prédit que
32f
Identi…cation de modèles proie-prédateur (thèse de doctorat)
C. Jost
la concentration du substrat à la sortie N ? ne devrait dépendre que de D, max et Ks ,
mais pas de la concentration du substrat à l’entrée N0 . Toutefois, on a rapidement observé que dans de nombreuses expériences la concentration du substrat à la sortie était
proportionnelle à celle du substrat à l’entrée. Contois (1959) a tenté de rendre compte
de cette observation en faisant l’hypothèse (et en apportant des données expérimentales
allant dans ce sens) que la constante de demi-saturation était proportionnelle à la concentration du substrat à l’entrée. En combinant cette observation avec le principe de
conservation de la masse (voir chapitre 6 pour plus de détails), il a dérivé la forme
(N, P ) =
um N
BP + N
(avec les constantes um et B), qui est une forme particulière de modèle ratio-dépendant.
Les articles qui ont testé cette fonction par rapport à celle de Monod ou d’autres
fonctions (voir Table 6.1 pour des exemples) peuvent être classés en deux catégories :
des expériences en chémostat avec des concentrations du substrat à l’entrée et des
taux de dilution variables, où la concentration du substrat est mesurée à la sortie,
des mesures directes de croissance, avec des concentrations variables du substrat
et des organismes, puis une comparaison qualitative et quantitative de ces mesures
avec diverses fonctions de croissance.
La première catégorie a montré que la prédiction de Monod est correcte tant que
l’on travaille avec une souche unique d’organismes croissant sur un substrat pur. Toute
modi…cation de ces conditions (substrat mixte ou souches/espèces d’organismes multiples) mène à ce que la concentration du substrat à la sortie soit proportionnelle à celle à
l’entrée. On peut interpréter ce résultat en disant que l’isocline du consommateur dans
un chémostat (Figure 6.2) doit avoir une pente positive non in…nie. La seconde catégorie
a montré que la fonction de croissance est une fonction décroissante de la densité des
consommateurs,
d(N, P )
< 0,
dP
sous les mêmes conditions que celles mentionnées pour la première catégorie (substrat mixte ou multiples souches/espèces d’organismes). Cependant, ces résultats sont
plus précis en ce qu’ils identi…ent la prédateur-dépendance dans la fonction de croissance comme la cause probable de la pente de l’isocline, et non une mortalité densitédépendante ainsi que l’ont proposé d’autres chercheurs. Si tous ces résulats suggèrent
que l’on doit s’attendre à ce que, dans les systèmes naturels, la pente de l’isocline soit
non in…nie, ceci est également prédit par toute fonction de croissance dépendante du
consommateur, et non pas uniquement par l’équation de Contois. De même qu’en écologie (section 1.2.5f), la comparaison entre les modèles de Monod et de Contois ne se fait
pas sur un pied d’égalité. Heureusement, il y a également eu des comparaisons quantitatives entre des données et les modèles, fondées sur des critères statistiques tels que
le meilleur ajustement. Celles-ci ont montré que, dans de nombreux cas (traitement
C. Jost
Identi…cation de modèles proie-prédateur (thèse de doctorat)
33f
des eaux usées, processus de fermentation), la fonction de Contois s’ajuste mieux que
celle de Monod, mais que souvent l’identi…cation est moins claire ou d’autres formes
fonctionnelles s’ajustent tout aussi bien que l’équation de Contois.
On peut tirer plusieurs enseignements des travaux faits en microbiologie :
Les fonctions proie-dépendantes (ou substrat-dépendantes) telles que celle proposée par Monod semblent fournir une bonne description du processus de consommation d’un substrat pur par une espèce isolée.
Lorsque le système devient plus complexe (plusieurs espèces, hétérogénéités spatiales ou temporelles), des fonctions prédateur-dépendantes (de n’importe quelle
forme, pas forcément ratio-dépendantes) sont plus adéquates.
Si la sélection de modèle par ajustement est déjà ambiguë pour des données de
microbiologie, alors on ne peut s’attendre à des résultats plus concluants lorsqu’on
travaille avec des données de dynamique de populations de terrain, à moins de
développer des méthodes d’analyse plus …nes comme celles proposées aux chapitres
7 et 8.
2.5f
Comparaison quantitative des modèles
Cette section résume les idées et les résultats de l’analyse développée dans les chapitres
7 et 8.
J’ai émis plus haut l’opinion selon laquelle les comparaisons existantes entre le modèle proie-dépendent et le modèle ratio-dépendant ne les mettent pas sur un pied d’égalité. L’analyse du chapitre 4 est déjà plus proche de cet objectif lorsqu’elle compare les
prédictions de chaque modèle avec les dynamiques observées dans les lacs d’eau douce.
Dans le présent chapitre, je vais comparer les deux modèles (2.1) quantitativement sur
la base des dynamiques temporelles qu’ils prédisent, en les ajustant directement à des
séries chronologiques de systèmes proie-prédateur. Le concept de la vraisemblance (Edwards 1992) fournit le cadre statistique permettant la comparaison des deux modèles
sur un pied d’égalité. L’idée est d’identi…er la vraisemblance entre modèles et données
à la probabilité d’obtenir les données étant donné le modèle (avec des paramètres qui
maximisent cette probabilité, le maximum de vraisemblance). Cette identi…cation est
intuitive (voir Press et al. (1992) pour quelques commentaires) mais reste très utile pour
la sélection de modèles. L’estimation par le maximum de vraisemblance est caractérisée
par plusieurs traits intéressants : l’estimation est consistante et il s’agit de l’estimateur
présentant la plus faible variance possible (Huet et al. 1992). Toutefois, la distribution
des erreurs doit être connue parfaitement pour que ces propriétés restent valables. L’estimation par le maximum de vraisemblance n’est pas très robuste par rapport à une
incertitude sur cette distribution.
Le concept de vraisemblance peut aussi être étendu à la comparaison de modèles
comportant un nombre différent de paramètres, grâce à l’utilisation de critères dits
‘d’information’. Le plus connu est le critère d’information d’Akaike AIC (Akaike 1973;
34f
Identi…cation de modèles proie-prédateur (thèse de doctorat)
C. Jost
Bozdogan 1987), mais il en existe d’autres, tels que le critère d’information de Bayes
BIC, de Mallow Cp (voir Hilborn & Mangel (1997) pour des références et une discussion de ces trois critères) ou le ‘consistent Akaike information criterion’ CAIC (Jones
1993). Cependant, dans la mesure où j’ai travaillé sur des modèles d’égale complexité,
je mentionne ces critères pour mémoire mais ils ne sont pas pertinents pour la présente
analyse.
Contrairement aux expériences à l’équilibre de Arditi & Saïah (1992), nous approchons la question de la sélection de modèle du point de vue de la dynamique. Nous
examinons si des données sur des séries temporelles de systèmes proie-prédateur peuvent révéler l’existence ou non de prédateur-dépendance dans la réponse fonctionnelle.
La comparaison sera faite en ajustant les modèles proie-dépendant et ratio-dépendant,
qui sont de complexité égale (2f.1), à des séries chronologiques, et nous utiliserons un
critère de meilleur ajustement (vraisemblance).
Les séries temporelles que j’utilise dans les analyses sont soit extraites de la littérature (protozoaires, arthropodes et systèmes proie-prédateur planctoniques) soit des
données originales sur le phyto- et le zooplancton dans le Lac Léman. Elles sont caractériées, comme la plupart des données en écologie, par un faible nombre (10 - 50
triplets de données {temps, proie, prédateur}) et par des erreurs assez grandes avec
des coefficients de variation pouvant aller jusqu’à 50%. Aussi la plus grande prudence
s’impose-t-elle dans les déductions quant à la ‘justesse’ biologique du formalisme utilisé
pour décrire les processus biologiques, même lorsque certains modèles s’ajustent très
bien. Ce problème a été illustré de façon convaincante dans le cas de la croissance
logistique (introduite par Verhulst (1838)) très tôt dans l’histoire de la biologie des
populations mathématique, en l’occurrence en 1939 par W. Feller. Celui-ci a choisi
deux autres modèles mathématiques ayant une forme en ‘S’ et qui ont le même nombre
de paramètres et a montré que les deux formes s’ajustaient tout aussi bien (ou même
mieux) aux données qui, à l’époque, étaient considérées comme les meilleures preuves
que la croissance logistique était de même nature qu’une loi physique. Feller a poursuivi en montrant que ces modèles prédisent également la même réponse à un taux de
prélèvement par individu constant que le modèle logistique.
Plus récemment, Simons & Lam (1980) ont montré que les mêmes précautions s’imposent dans le cas de modèles plus complexes tels que ceux décrivant la dynamique du
phosphore dans les lacs de grande taille. Ils ont indiqué que divers jeux de paramètres
(avec ou sans saisonnalité imposée) donnent des ajustements de même qualité à une
saison de données, et que par conséquent le pouvoir prédictif d’un tel modèle (ajusté)
est extrêmement limité.
Ces deux exemples montrent amplement que le fait d’obtenir un bon ajustement
d’un modèle à des données est loin d’être une preuve de la ‘justesse’ biologique de la
description mathématique du processus que l’on a choisie. Autrement dit, ainsi que l’a
formulé Cale et al. (1989) , ‘de multiples con…gurations de processus peuvent produire
un même pattern’. Cependant, il existe une solution, consistant à suivre le conseil de
May (1989) de ‘générer des pseudo-données dans des mondes imaginaires dont les règles
sont connues, puis de tester l’efficacité des méthodes conventionnelles pour révéler ces
règles connues’. En d’autres termes, je vais générer des données sur des systèmes proieprédateur arti…ciels, leur appliquer les techniques de sélection de modèles, et véri…er si
C. Jost
Identi…cation de modèles proie-prédateur (thèse de doctorat)
35f
elles permettent d’identi…er le bon modèle. Ce n’est qu’après avoir évalué les limites
de l’identi…cation de modèles dans cette simulation que je tenterai d’analyser des séries
temporelles écologiques réelles. Cette analyse est résumée dans la section 2.5.4f.
Cette même approche a été utilisée par Carpenter et al. (1994) pour des données de
plancton, mais les auteurs ont utilisé un modèle discret où les pas de temps étaient les
intervalles de temps (arbitraires) entre les mesures. Ils ont également ajusté des séries
temporelles sur plusieurs années, supposant par conséquent que des paramètres tels que
la capacité de charge ou le taux de mortalité ont la même valeur d’une année sur l’autre.
Je ré-analyserai leurs données en ajustant nos modèles aux données d’une seule saison,
en particulier à la période entre le printemps et l’automne où les interactions entre
populations, plutôt que des contraintes physiques telles que la température ou la lumière,
pourraient être les forces déterminant principalement les dynamiques (Sommer et al.
1986). J’ajusterai également les modèles aux séries temporelles sur des protozoaires
extraites de la littérature car elles présentent une variabilité beaucoup plus faible que
les données sur le plancton dans des lacs et parce qu’il y a de nombreuses similarités
entre les protozoaires et le plancton (un milieu ‡uide, des générations chevauchantes,
des effectifs importants, etc.). Il y a également plus de modèles écologiques incluant une
boucle microbienne (‘microbial loop’, par ex. Fasham et al. 1990, Azam et al. 1983,
Stone & Berman 1993) fondés sur une quantité importante de données empiriques sur
le rôle écologique important de cette faune microbienne (Berman 1990; Fenchel 1988;
Sherr & Sherr 1991). Il sera intéressant de voir comment des modèles typiquement
écologiques s’ajustent à des données concernant de tels organismes.
En…n, ce travail présente également des différences par rapport à l’approche présentée par Harrison (1995). L’auteur a utilisé une série temporelle classique de modèle
proie-prédateur extraite de la littérature (Luckinbill 1973) et l’a ajustée à 11 modèles
proie-prédateur différents, en essayant de déterminer les processus clé nécessaires pour
obtenir un bon ajustement qualitatif et quantitatif. Deux problèmes se posent dans ce
travail : (1) l’auteur a implicitement supposé qu’un meilleur ajustement est le signe
d’une meilleure description des processus individuels et (2) en l’absence d’informations
sur l’importance de l’erreur d’observation, aucun test statistique tenant compte du nombre de paramètres des différents modèles n’a pu être utilisé dans la comparaison, par ex.
le test du rapport des vraisemblances (Hilborn & Mangel 1997) ou le critère d’information d’Akaike (AIC, Akaike 1973). Je montrerai que, déjà avec une erreur d’observation
modeste dans les données, l’hypothèse (1) peut être fausse et que des travaux détaillés de
simulation doivent être faits avant une telle étude. Il n’est donc pas surprenant que dans
l’analyse de Harrison (1995) le modèle avec le plus grand nombre de paramètres soit
celui qui s’ajuste le mieux. Tandis que ce type d’analyse peut fournir des informations
sur quelques détails des processus de systèmes spéci…ques (par ex. l’importance du retard entre consommation de Paramecium aurelia et reproduction de Didinium nasutum
dans cette analyse), il ne permet pas de savoir si le fait d’ajouter ces détails améliore
signi…cativement l’ajustement aux données. J’éviterai ce type d’erreur de raisonnement
en travaillant uniquement avec des modèles contenant le même nombre de paramètres.
36f
2.5.1f
Identi…cation de modèles proie-prédateur (thèse de doctorat)
C. Jost
Une remarque sur le critère de sélection de modèle
Dans cette thèse, la sélection de modèle est toujours fondée sur la vraisemblance (meilleur
ajustement), au moins lorsque je travaille avec des données réelles. Une autre approche
a été choisie par Bilardello et al. (1993), qui a utilisé la région de con…ance jointe
linéarisée (‘joint linearised con…dence region’) comme critère de comparaison. Cette
région est minimale si le déterminant de J ()T J () est maximal (Hosten 1974), J ()
étant la matrice de sensibilité des variables d’état par rapport au vecteur de paramètres
(connu également sous le nom de matrice d’information de Fisher, Huet et al. 1992).
Je n’ai pas testé ce critère, mais il est évident qu’il aiderait à la détection de problèmes
liés à la surparamétrisation (qui rend cette matrice proche de devenir singulière). J’ai
rencontré de tels problèmes avec les niveaux d’erreurs les plus élevés dans l’analyse ‘in
silico’ (2.5.4f). La convergence était parfois très lente, même proche de l’optimum, et
parfois plusieurs jeux de paramètres donnaient une qualité d’ajustement équivalente.
Les deux phénomènes indiquent qu’un trop grand nombre de paramètres ont été ajustés
aux données. Un autre critère bien connu est l’écart-type des résidus (Carpenter et al.
1994).
Tous ces critères peuvent soulever le problème du fait que la sélection de modèles
se base sur une quantité qui n’a pas été minimisée en ajustant le modèle aux données.
J’analyserai deux critères de ce type proposés par Carpenter et al. (1994) dans le
chapitre 7. Souvent, le modèle sélectionné donnait visuellement un ajustement moins
bon que le modèle rejeté (voir section 2.5.4f). Le fait de n’utiliser que le concept de la
vraisemblance pour les régressions et la sélection de modèles évite ce problème, et il est
statistiquement bien fondé. C’est la raison pour laquelle je ne me suis servi que de ce
concept dans l’analyse des données réelles (chapitre 8).
2.5.2f
Les erreurs qui gouvernent notre monde modélisateur
J’ai mentionné dans la section précédente que ma sélection de modèle se base sur une
approche fondée sur la vraisemblance. Pour formuler cette vraisemblance, je dois formuler un modèle stochastique, i.e., un modèle qui explique comment des effets aléatoires
entrent dans les données auxquelles je veux ajuster mes modèles dynamiques. Ces modèles eux-mêmes sont déterministes. Ils ont la forme d’équations différentielles ordinaires
(EDO) qui, pour des valeurs données des paramètres, donnent des résultats complètement prédictibles. Ils forment le noyau du modèle statistique auquels on ajoute des
stochasticités.
Un premier type de stochasticité est une pure erreur d’observation ou de mesure:
l’effectif de la population a une valeur précise, mais chaque estimation n’est qu’une
réalisation d’un nombre aléatoire selon une certaine distribution. Si nous avons des
données qui ne contiennent que des erreurs d’observation (la dynamique de la population
étant déterministe), nous ajustons le modèle en ajustant la trajectoire entière à la série
chronologique (solution d’une EDO où les conditions initiales sont traitées comme les
autres paramètres à estimer). Notons que dans ce traitement nous considérons que les
mesures de temps sont exactes, sans erreur. Ce type d’ajustement est souvent appelé
‘trajectory …tting’. Voir Figure 8.1a pour une illustration de ce type d’ajustement.
C. Jost
Identi…cation de modèles proie-prédateur (thèse de doctorat)
37f
Un deuxième type de stochasticité est l’erreur de processus : la dynamique d’une
population n’est pas entièrement dé…nie par sa composante déterministe, mais aussi
par une in‡uence aléatoire provenant soit de l’environnement soit de la population
elle-même. Dans le contexte d’une EDO cette erreur de processus crée une équation
différentielle ordinaire stochastique (EDOS). Avec un système de ce type, la trajectoire
déterministe du modèle diverge de plus en plus par rapport à l’abondance réelle de la
population. Plus l’horizon de prédiction est éloigné, pire est la qualité de la prédiction.
L’abondance de la population après un certain intervalle de temps …xé avec une condition
initiale (qui est connue si nous mesurons l’abondance sans erreur d’observation) est
alors une réalisation d’un processus stochastique. Pour relier dans ce cas le problème de
l’estimation de paramètres (et de l’estimation du meilleur ajustement) à un problème
de régression dans le sens classique du terme, nous devons faire l’hypothèse que la
distribution de la population après cet intervalle de temps, étant donnée la condition
initiale, est connue. L’horizon de prédiction est généralement dé…ni par la longueur du
pas de temps dans la série chronologique, mais il est également possible de prédire s > 1
pas de temps à l’avance. Ce type d’ajustement est souvent appelé ‘s-step-ahead-…tting’.
Voir Figure 8.1b pour une illustration de ce type d’ajustement.
Solow (1995) distingue un troisième type d’erreur, appelée erreur de paramètre. Ceci
est un cas particulier de l’erreur de processus, qui est due à la nature stochastique d’un
(ou plusieurs) paramètres du modèle. Solow démontre que ce type d’erreur peut rendre
l’erreur globale non stationnaire, dépendant de la fonction de densité et de l’encastrement algébrique de ce paramètre. Dans mon problème, je ne voyais pas de possibilité
pour distinguer entre erreur de paramètre et erreur de processus en général, ainsi qu’elle
est modélisée avec une EDOS. Outre cela, la structure algébrique des équations (2f.1)
rend difficile de savoir comment un erreur de paramètre se traduit en erreur de processus en général, une information nécessaire pour adapter la procédure de régression. J’ai
donc seulement considéré l’erreur de processus sous la forme d’une EDOS, suivant le
traitement standard des erreurs dans les séries chronologiques (Hilborn & Mangel 1997;
Pascual & Kareiva 1996; Dennis et al. 1995).
La théorie statistique a développé principalement des méthodes de régression tenant compte d’un seul type d’erreur, le problème de considérer conjointement erreur
de processus et erreur d’observation étant beaucoup plus difficile (voir discussion à la
section 3.2.3f). Si le chercheur sait choisir son plan d’éxpérience d’une manière qui
permet d’éliminer (ou du moins de réduire à des valeurs négligeables) l’un des deux
types d’erreur (par ex., en créant un environnement constant dans le laboratoire et en
travaillant avec des tailles de populations qui rendent la stochasticité démographique
négligeable, ou en développant une méthode d’échantillonnage sans erreur d’observation), alors aucun problème ne se pose en formulant le modèle stochastique. Cependant, ceci est difficile à atteindre, surtout pour des études de terrain. La majorité des
séries chronologiques disponibles dans la littérature ou sur les expériences propres d’un
chercheur contiennent les deux types d’erreur. Pascual & Kareiva (1996) arrivent alors
à la conclusion qu’un chercheur doit se décider (avec une part d’arbitraire) pour la modélisation d’une seule erreur. En raison de la pertinence de ces travaux, je vais souvent
employer leur terminologie, en appelant le cas d’ajustement de la trajectoire entière
‘observation error …t’ et le cas de l’ajustement s pas en avant ‘process error …t’.
38f
Identi…cation de modèles proie-prédateur (thèse de doctorat)
C. Jost
Le choix particulier de l’un ou de l’autre type d’erreur peut in‡uencer profondément
le résultat d’une sélection de modèle. Cependant, la sensibilité de la sélection de modèle
pour un type d’erreur particulier peut être étudiée par une analyse de simulation, comme
proposé par May (1989). Ceci sera le sujet de la section suivante. Je discute et étudie là
aussi des fonctions d’erreur proposées par Clutton-Brock (1967) (et appliquées par Carpenter et al. (1994)) qui semblent tenir compte des deux types d’erreurs sans augmenter
trop fortement les coûts numériques. Cependant, ces fonctions d’erreur ne s’intègrent
pas dans le concept de vraisemblance, soulevant le problème du choix d’un critère pour
la sélection de modèle. La section suivante va illustrer quelques conséquences de ce problème. Une autre approche statistique prenant en compte les deux types d’erreurs est
l’approche des erreurs dans les variables (‘errors-in-variables’) qui sera discutée brièvement dans la section 3.2.3f. Je n’ai pas utilisé cette méthode parce que les coûts de
calcul sont beaucoup plus élevés, que la méthode n’en est encore qu’à ses débuts, et
qu’elle requiert une information indépendante soit sur l’erreur de processus soit sur l’erreur d’observation. Cependant, cette méthode s’intègre parfaitement dans le concept
de vraisemblance et semble prometteuse pour de futures recherches.
Notons que je fais l’hypothèse, pour l’‘observation-error-…t’ et le ‘process-error-…t’
que les erreurs sont stationnaires et normales sur une échelle logarithmique (coefficient
de variation CV constant). Avec cette hypothèse, la méthode du maximum de vraisemblance est équivalente à celle des moindres carrées (Press et al. 1992), et quand j’utilise
les termes de ‘meilleur ajustement’ ou de ‘vraisemblance’, le lecteur s’en fait plus facilement une idée en pensant aux moindres carrées. Pourtant, la taille ce cette erreur
(CV ) n’est connue qu’approximativement pour des données réelles. De plus, ces séries
chronologiques contiennent les deux types de stochasticité, et négliger l’une d’entre
elles n’est qu’une nécessité statistique. Ceci veut dire que les vraisemblances calculées
n’auront pas une interprétation absolue, et que les critères d’information mentionnés
précédemment ne peuvent pas être utilisés. Pour cette raison, la séléction ne peut se
faire que pour des modèles de même complexité (même nombre de paramètres).
2.5.3f
Implémentation des algorithmes d’ajustement
Avant de discuter les résultats des ajustements eux-mêmes, je dirai quelques mots sur
l’implémentation des algorithmes d’ajustement. La plupart des algorithmes ont tout
c a…n de savoir s’ils étaient opérationnels.
d’abord été développés dans Mathematica
Cependant, les calculs numériques dans Mathematica sont trop lents pour permettre
l’analyse d’un grand nombre de séries temporelles. Il n’existe pas beaucoup de logiciels
permettant l’ajustement d’équations différentielles à des données que l’on puisse adapter
à ses propres besoins. L’ajustement et la visualisation du modèle ajusté se font dans la
plupart des logiciels en deux étapes, ce qui ralentit encore plus le processus d’ajustement.
C’est la raison pour laquelle j’ai implémenté tous les algorithmes en C++, en utilisant
la boîte à outils MacIntosh pour visualiser le modèle ajusté en temps réel. Ce contrôle
visuel immédiat s’est avéré indispensable pour l’analyse d’un grand nombre de jeux
de données. Les algorithmes numériques que j’ai utilisés se basent sur le code livré
avec le livre ’Numerical Recipes in C’ (Press et al. 1992). Le logiciel peut être obtenu
directement auprès de l’auteur (il nécessite l’emploi d’un PowerMacIntosh).
C. Jost
Identi…cation de modèles proie-prédateur (thèse de doctorat)
2.5.4f
Une approche ‘in silico’ de la sélection de modèle
39f
L’objectif de l’analyse décrite dans le chapitre7 est d’étudier numériquement comment
la présence des deux types d’erreur, erreur de processus et erreur d’observation, in‡uence la sélection de modèle par meilleur ajustement. Pour cela, j’ai créé des séries
chronologiques arti…cielles qui contiennent les deux erreurs. Le noyau déterministe est
inspiré par les dynamiques décrites dans le modèle PEG (section 2.2f), c’est à dire
qu’il est constitué de dynamiques qui convergent vers un équilibre stable après une ou
plusieurs oscillations de grande amplitude. Dans une première étape, j’ai calculé par
une procédure aléatoire des paramètres et des conditions initiales pour le modèle proiedépendant et pour le modèle ratio-dépendant (2f.1). Les critères pour accepter un tel
jeu de paramètres aléatoires sont d’une part la conformité aux dynamiques décrites cidessus, et d’autre part des nécessités pratiques telles que le fait que l’équilibre de la proie
et du prédateur diffèrent par un facteur inférieur à 100 et que les dynamiques du prédateur varient au minimum par un facteur 5. Pour chaque modèle, 20 jeux de données
ont été créés de cette façon. Par analogie avec des expériences en écologie, j’ai simulé 5
séries chronologiques ‘réplicats’ pour chaque jeu de données, en ajoutant une erreur de
processus dans le cadre d’une équation différentielle stochastique, en ‘échantillonant’ la
série chronologique résultante à 20 intervalles de temps équidistants (ceci est également
inspiré du modèle PEG, 20 étant le nombre typique d’échantillons de plancton prélevés
durant une saison) et en ajoutant a ces échantillons une erreur d’observation. L’erreur
de processus est générée dans le cadre d’un processus stochastique standard (voir détails dans la section 7.3.1) avec une composante stochastique qui est proportionnelle à
l’abondance de la population actuelle, simulant ainsi une erreur de type multiplicatif,
qui est approximativement log-normale. L’erreur d’observation est de type log-normal
avec un coefficient de variation constant. Des séries chronologiques avec des erreurs
faibles (comparables à celles de systèmes protozoaires en conditions de laboratoire) et
avec des erreurs fortes (comparables à celles dans des mésocosmes aquatiques, peut-être
même dans des lacs entiers) ont été créées par cette procédure.
Ensuite, chaque modèle (proie-dépendant et ratio-dépendant) a été ajusté à chaque
série chronologique avec quatre méthodes (ou fonctions d’erreurs). La méthode A fait
l’hypothèse qu’il y a seulement une erreur d’observation de type log-normal, ‘observation
error …t’. Ceci correspond au critère des moindres carrés en échelle logarithmique. La
méthode B suppose qu’il y a seulement une erreur de processus de type lognormal,
‘process error …t’. Dans ce cas, le maximum de vraisemblance correspond au critère
des moindres carrés conditionnels en échelle logarithmique. L’horizon de prédiction s
est choisi de sorte que l’autocorrélation dans la série chronologique soit inférieure à
0.5 (Ellner & Turchin 1995). La raison de ce choix est qu’il faut faire l’ajustement
sur un intervalle de temps suffisamment long pour que des effets non-linéaires soient
aussi importants (ou plus importants) que les effets linéaires. En fait, ce critère pour
la détermination de l’horizon de prédiction est inspiré de Ellner & Turchin (1995),
et justi…é dans cet article-là comme un résultat empirique. Les méthodes C et D
sont inspirées de Clutton-Brock (1967), de la façon dont elles ont été présentées par
Carpenter et al. (1994). C correspond à des moindres carrés conditionnels pondérés, où
les poids dépendent de la sensibilité de la prédiction aux conditions initiales, qui sont
les observations s pas de temps plus tôt, et qui contiennent une erreur d’observation
40f
Identi…cation de modèles proie-prédateur (thèse de doctorat)
C. Jost
(voir section 7.3.2 pour plus de détails). L’erreur d’observation est supposée connue
indépendamment de l’erreur de processus. La méthode D est dérivée de C, mais elle
ressemble plus au log de l’opposé de la vraisemblance de manière à ce que le facteur
d’échelle (’scaling factor’) de la distribution normale soit conservé (voir Carpenter et al.
1994), de sorte que des poids plus élevés ne diminuent pas seulement l’importance du
résidu mais ajoutent aussi un facteur de pénalisation à la fonction d’erreur. Notons que
ces fonctions de perte C et D (Carpenter et al. 1994) n’entrent pas dans le concept de
la vraisemblance. La sélection de modèle se base donc sur un autre critère. J’ai choisi
la somme des carrés des résidus (calculée en échelle log ) comme critère de sélection.
Les séries chronologiques avec des erreurs faibles (erreur d’observation et erreur de
processus) ont été identi…ées d’une manière sûre (moins de 5% d’identi…cations fausses)
par les méthodes A, B et D. La méthode C a donné près de 15% d’identi…cations
erronnées, et a donc été rejetée en tant que méthode utile pour la sélection de modèle
(voir Figure 7.3).
Les séries chronologiques avec des erreurs élevées ont été identi…ées de la façon la
plus satisfaisante par la méthode A (‘observation error …t’), avec moins de 5% d’identi…ciations erronées. Les méthodes B et D ont toutes deux donné 10% de mauvaises
identi…cations, mais B s’est avérée préférable dans le sens où, si l’on exige une différence
minimale de 5% entre les valeurs du critère de sélection, alors cette méthode a identi…é
plus de 95% des séries chronologiques correctement (Figure 7.3). Aucun seuil de ce
genre n’a pu être trouvé pour la méthode D.
Les problèmes avec les méthodes C et D sont de deux natures différentes. Premièrement, les algorithmes ont souvent maximisé la dépendance aux conditions initiales
(créant des poids importants) plutôt que de minimiser les résidus. Deuxièmement, le
critère de sélection de modèle a souvent augmenté pendant la minimisation du critère
d’ajustement (la fonction de perte), et la valeur minimum de la fonction de perte a pu
être arbitrairement loin de son propre minimum. La Figure 2f.5 illustre ces problèmes
pour un jeu de données particulier. J’ai donc conclu que les deux méthodes C et D
sont de faible utilité pour la sélection de modèle.
Le résultat le plus intéressant est que les deux méthodes A et B n’identi…ent le
modèle faux que dans 1% des séries chronologiques analysées. L’identi…cation la plus
…able est donc d’appliquer plusieurs méthodes (par ex. A et B ) et d’accepter le résultat
de la sélection de modèles seulement si toutes les méthodes ont identi…é le même modèle.
Un résultat supplémentaire de cette analyse est la qualité des paramètres estimés par
ajustement d’une EDO à des séries chronologiques. Pour chaque ajustement d’un modèle à une série chronologique donnée (créée avec le même modèle), j’ai calculé le ratio du
paramètre estimé divisé par le paramètre original (avec lequel la série chronologique a été
créée). La distribution cumulée de ces ratios (après ajustement aux séries chronologiques
avec des erreurs élevées) est présentée dans la Figure 2f.6. Si la courbe de cette distribution cumulée passe par le point (1.0, 0.5) l’espérance de la médiane est correcte.
Plus cette courbe est raide, moins les paramètres estimés sont dispersés. Nous voyons
que seuls les paramètres r et K sont identi…és d’une manière satisfaisante pour les deux
modèles, tous les autres paramètres étant dispersés largement et la médiane étant souvent éloignée de 1 (où elle devrait être). Toutes les fonctions d’erreur analysées ont
C. Jost
Identi…cation de modèles proie-prédateur (thèse de doctorat)
25
41f
150
20
100
15
10
50
5
200
400
600
800
200
400
600
800
Figure 2f.5: Illustration des problèmes observés avec la méthode de régression
C. Les graphiques montrent dans l’espace de phase (N : proie, P : prédateur,
IsoN : isocline de la proie, IsoP : isocline du prédateur) comment le modèle
proie-dépendent a été ajusté par la méthode B ((a), ajustement avec erreur de
processus) et par méthode C ((b), moindres carrés conditionnels pondérés) à une
série chronologique stochastique créée avec le modèle proie-dépendant. La méthode
C a, en fait, maximisé la dépendance aux conditions initiales au lieu de minimiser
les résidus, résultant en un ajustement qui est évidemment faux. Des problèmes
similaires ont été observés avec la méthode D. Notons que les axes du prédateur
ont des échelles différentes en (a) et en (b).
donné d’aussi mauvais résultats.
La qualité du jeu de paramètres entier (par modèle et par série chronologique) peut
être visualisée d’une manière similaire en travaillant avec la valeur propre dominante
(stabilité locale) de l’équilibre non-trivial. Ces valeurs propres ont été calculé algébriquement pour éviter les erreurs numériques. A présent, on observe de fortes différences entre
les méthodes. La Figure 7.5 montre les distributions cumulées des ratios de la valeur
propre dominante réelle divisée par la valeur propre dominante estimée pour toutes les
fonctions d’erreur et pour chaque modèle. La raideur de chaque courbe est approximativement la même, donc chaque méthode donne une estimation de la stabilité locale
dispersée d’une manière équivalente (toutefois, il semble y avoir moins de variations
avec le modèle ratio-dépendent). Cependant, selon le deuxième critère - la déviation
par rapport à la médiane espérée -, on voit que la méthode A (‘observation error …t’,
ou ajustement avec erreur d’observation) est globalement la meilleure, suivie par la
méthode D. La méthode B (‘process error …t’, ou ajustement avec erreur de processus)
surestime la stabilité locale.
Conclusions de l’approche ‘in silico’
Ce qu’il faut retenir de cette analyse est que l’identi…cation de modèle est en principe
possible avec les deux niveaux d’erreur et le type de série chronologiques étudiés (dynamique convergeant vers une équilibre stable après une ou plusieurs oscillations, commençant à des conditions initiales faibles et échantillonée 20 fois). L’ajustement avec
erreur d’observation est légèrement plus …able que l’ajustement avec erreur de proces-
42f
Identi…cation de modèles proie-prédateur (thèse de doctorat)
C. Jost
Prey-dependent data
1
1
1
r
0.6
K
1.4
1
Cumulative distribution
1
0.6
1.4
1
1
1.4
1
e/ c
0.6
1.4
1
1
h
0.6
a
µ
e
0.6
1
1.4
e/ c
0.6
1
1.4
e/ c
Ratio-dependent data
1
1
1
K
r
0.6
1.4
1
1
0.6
1
e/ c
1.4
0.6
1.4
1
1
1
h
0.6
1.4
1
e
0.6
1
e/ c
µ
1.4
0.6
1.4
1
e/ c
Figure 2f.6: Analyse de la qualité de l’estimation après ajustement aux séries chronologiques à erreurs élevées, par paramètre et par modèle : distribution cumulative
des rapports e /c (e étant le paramètre estimé, c le paramètre original qui a créé
les données) pour toutes les fonctions d’erreurs. — — — : ajustement avec erreur
d’observation A ; ——— : ajustement avec erreur de processus B; – – – – :
moindres carrés pondérés (méthode D, voir texte).
sus, mais la meilleure identi…cation est obtenue en identi…ant avec les deux méthodes et
en acceptant une sélection seulement si les deux méthodes identi…ent le même modèle.
La qualité des paramètres estimés est généralement très mauvaise et ces estimations ne
devraient être utilisées qu’avec beaucoup de précautions. La méthode d’ajustement des
équations différentielles ordinaires est donc plus appropiée pour la sélection de modèles
que pour l’estimation de paramètres.
C. Jost
2.5.5f
Identi…cation de modèles proie-prédateur (thèse de doctorat)
43f
L’analyse ‘in vivo’
Comme on l’a vu dans la section précédente, la sélection de modèle est plus …able si
plusieurs critères de meilleur ajustement, sur la base de plusieurs modèles statistiques
différents, sélectionnent le même modèle. Les modèles statistiques les mieux connus proposent les méthodes d’ajustement avec erreur d’observation et d’ajustement avec erreur
de processus. En conséquence, je n’analyserai que des séries chronologiques permettant
les deux types d’ajustement. Ceci est particulièrement contraignant pour l’ajustement
avec erreur d’observation, car plus les séries chronologiques sont longues, plus la probabilité de pouvoir raisonnablement ajuster une trajectoire entière est faible. Bien sûr, on
pourrait diviser des séries chronologiques longues en petits ’morceaux’ et ajuster les trajectoire à chacun d’eux, mais il n’existe pas de méthode statistique standardisée sur la
façon de faire correctement cette division. Pour éviter toute ambiguïté liée à des méthodes statistiques non-standardisées, je n’utiliserai que des séries chronologiques qui peuvent être ajustées raisonnablement à des trajectoires entières. Ces séries chronologiques
vont de cultures en batch de protozoaires (Gause 1935; Gause et al. 1936; Luckinbill
1973; Veilleux 1979; Flynn & Davidson 1993; Wilhelm 1993), via des systèmes proieprédateur de laboratoire spatialement plus complexes (Huffaker 1958; Huffaker et al.
1963), à des systèmes planctonique très complexes de lacs d’eau douce de la zone tempérée (Carpenter et al. 1994; CIPEL 1995). Les données planctoniques de Carpenter
et al. (1994) consistent en phytoplancton comestible et zooplancton herbivore de deux
lacs Nord-américains, de 1984 à 1990 (Paul Lake et Tuesday Lake). Le phytoplancton
a été dé…ni par les auteurs comme étant l’ensemble des organismes photosynthétiques
ayant un biovolume inférieur à celui d’une sphère de diamètre 30 m. Pour le Lac Léman je disposais du phtyoplancton comestible (dé…ni comme l’ensemble des organismes
présentant une longueur inférieure à 50m et un biovolume inférieur à 104 m 3 ) et du
phytoplancton total. J’ai donc ajusté les modèles à deux systèmes : phytoplancton
comestible – zooplancton herbivore, et phytoplancton total – zooplancton herbivore.
Quelles sont les différences entre ces données réelles et les données arti…cielles de
l’analyse ‘in silico’ ? Le caractère mulitplicatif de l’erreur d’observation (et probablement aussi de l’erreur de processus) a été con…rmé pour tous les systèmes. Cependant,
l’hypothèse de la normalité de ces erreurs (en données transformées en log ) pourrait
être trop forte, des ‘outliers’ (queues plus importantes, points plus éloignés que prévu
de la moyenne) sont possibles dans les données sur des arthropodes et dans celles sur le
plancton. J’ai donc également ajusté les modèles avec une fonction d’erreur basée sur la
somme des valeurs absolues des résidus en échelle logarithmique (distribution Laplacienne ou double exponentielle). Une autre différence est la taille absolue de l’erreur (CV ),
qui pourrait être plus élevée (surtout avec le phytoplancton) que dans les données arti…cielles. Dans ce cas, la différence de 5% entre les fonctions d’erreur après ajustement
des deux modèles (2f.1), qui a été calculée dans la section précédante pour assurer une
con…ance de 95% pour la sélection de modèle, n’est peut-être pas suffisamment élevée.
J’ai donc aussi fait un ‘bootstrapping résiduel’ (décrit par Efron & Tibshirani (1993),
voir détails dans la section 8.A) pour calculer l’intervalle de con…ance de l’erreur avec
chaque modèle et en appliquant un test t standard pour voir si l’un des modèles s’ajuste
signi…cativement mieux que l’autre (avec = 0.05). Le même algorithme fournit aussi
un ‘improved estimate of prediction error’ (IEPE, Efron & Tibshirani 1993) du modèle
44f
Identi…cation de modèles proie-prédateur (thèse de doctorat)
C. Jost
ajusté, qui peut être testé pour sa signi…cativité entre les deux modèles. En…n, une différence est qu’avec les données réelles, par dé…nition le processus à l’origine des données
est inconnu, ce qui signi…e que les deux modèles considérés peuvent être complètement
faux. Pour contrer ce problème je n’ai interprété que les ajustements dont la vraisemblance dépassait un certain seuil, estimé sur la base des vraisemblances obtenues après
ajustement des données arti…cielles. En somme, pour chaque type d’ajustement (ajustement avec erreur d’observation et ajustement avec erreur de processus) quatre critères
de sélection ont été appliqués, et une sélection de modèle n’a été considéré comme signi…cative que si tous les critères (avec une vraisemblance suffisamment élevée) avec les
deux type d’ajustement donnaient le même modèle.
Les résultats de sélection les plus signi…catifs ont été obtenus avec les données des
systèmes protozoaires. La majorité était soit plus proche de la proie-dépendence, soit
les échantillons étaient trop petits pour détecter la prédateur-dépendance d’une manière
sûre. Toutefois, il existe un système proie-prédateur, pour lequel quatre jeux de données
sont disponibles, qui présente une prédateur-dépendance signi…cative. Les prédateurs
dans ce système peuvent être fortement cannibales aux faibles densités de proies (comm.
pers. des auteurs). Bien qu’un tel cannibalisme n’ait pas été observé dans les conditions où les données analysées ont été produites, ceci suggère que les prédateurs sont
capables d’un fort degré d’interférence lorsqu’ils se rencontrent. Il est également possible que des hétérogénéités se soient développées entre les épisodes d’agitation de la
culture (toutes les 12 h). Les deux facteurs pourraient expliquer le fait que les données
soient, de façon très signi…cative, en faveur du modèle ratio-dépendant. A notre connaissance, ceci est le premier exemple d’un système de protozoaires avec une seule espèce
de proie et une seule espèce de prédateur qui présente cette forte prédateur-dépendance
(à comparer au chapitre 6). Cette exception illustre le fait que le modélisateur doit
bien connaître la biologie du système qu’il veut modéliser, et que des caractéristiques
telles que le cannibalisme potentiel peuvent indiquer qu’un modèle comportant de la
prédateur-dépendance est plus approprié (voir la section 8.5 pour des commentaires
plus développés et la mise en relation avec d’autres expériences).
Il ne semble pas non plus y avoir de prédateur-dépendance dans les données de
Huffaker sur des arthropodes. Dans la plupart des cas, le modèle proie-dépendant
donne le meilleur ajustement, et dans les deux cas de ‘process error …t’ où le modèle
ratio-dépendant donne le meilleur ajustement, ce dernier est qualitativement faux. Deux
autres aspects sont importants pour les ajustements à ces données : 1) quantitativement,
les modèles s’ajustent plutôt mal aux données, le système expérimental montrant de plus
grandes variations que nos modèles simples ne peuvent en reproduire et 2) ‘observation
error …t’ donne des ajustements corrects avec les deux modèles. Le premier point peut
s’expliquer par le dispositif expérimental de Huffaker, la nourriture pour les proies étant
dispersée dans une structure à 2 ou à 3 dimensions et la proie colonisant ses ressources
d’une façon relativement hétérogène. Un tel système de laboratoire est plus complexe,
du point de vue de sa structure, que les cultures de protozoaires en ‘batch’ du paragraphe
précédent. Le second point indique que les modèles utilisés peuvent toutefois être utilisés
pour l’analyse qualitative, et que seuls les résultats quantitatifs doivent être interprétés
avec précautions.
Ce sont les ajustements aux données de phyto- et de zooplancton qui sont les plus
C. Jost
Identi…cation de modèles proie-prédateur (thèse de doctorat)
45f
difficiles à interpréter. La conclusion la plus simple serait de dire soit que les données
sont trop ‘bruitées’ pour ce type d’identi…cation de modèle, soit que les deux modèles
sont trop simples pour représenter correctement la dynamique d’un lac. Un élément
en faveur du premier point est la nature qualitative de la régression de ‘process error …t’ (systèmes généralement stables ou fortement stables) qui pourraient indiquer
que la meilleure prédiction n’est pas celle obtenue par modélisation dynamique nonlinéaire mais plutôt celle obtenue en utilisant simplement une valeur moyenne des données comme prédiction. Le second point est probablement juste pour ‘observation error
…t’, mais pas nécessairement pour ‘process error …t’. En dépit de ces inconvénients, de
nombreuses identi…cations de modèle signi…catives sont obtenues avec des ajustement
avec ‘observation error …t’, indiquant qu’il existe des patterns de dynamique à long
terme. Ces ajustements signi…catifs concernent les deux types de modèles, proie- et
ratio-dépendant, avec des tendances pour certains lacs ou lorsqu’on modélise seulement
le phytoplancton comestible ou la totalité du phytoplancton. Cependant, ces tendances
ne sont pas suffisamment convaincantes pour donner quelque recommandation que ce
soit quant au choix du modèle le plus approprié. Selon Brett & Goldman (1997) le phytoplancton serait soumis à une forte in‡uence de bas en haut tandis que le zooplancton
serait plus sensible au contrôle par le haut. L’interaction phytoplancton-zooplancton
elle-même (qui est étudiée dans la présente thèse) est soumise aux deux types de forces,
ce qui pourrait également expliquer les ambiguïtés de la sélection de modèle.
Conclusions de l’analyse ‘in vivo’
Les systèmes avec une seule espèce de proie et une seule espèce prédatrice dans un
environnement homogène présentent généralement peu de prédateur-dépendance et sont
modélisés de façon plus satisfaisante par le modèle prédateur-dépendant. Toutefois, j’ai
trouvé au moins un système où le modèle ratio-dépendant s’ajuste mieux que le modèle
proie-dépendant pour l’ensemble des séries chronologiques disponibles. Aussi, avant
de modéliser de tels systèmes avec un modèle proie-dépendant, l’on devrait véri…er
si certains traits indiquant une forte prédateur-dépendance sont présents (par ex. le
cannibalisme potentiel). Avec les lacs, aucune conclusion en faveur de l’un ou de l’autre
modèle ne se dégage. D’une façon générale, lorsqu’on fait des prédictions avec un
modèle, il serait bon de procéder à des véri…cations en les comparant à celles d’un
autre modèle. La proie-dépendance et la ratio-dépendance offrent un cadre prometteur
pour développer de tels modèles alternatifs et pour détecter les prédictions ‘robustes’
(robustes au sens où elles sont insensibles à la prédateur-dépendance), et pour dé…nir
de futures directions de recherche lorsqu’aucune prédiction claire ne se dégage (càd
étudier l’ampleur de la prédateur-dépendance effectivement à l’oeuvre dans le système
ou paramétriser un modèle intermédiaire).
De Mazancourt et al. (1998) et Zheng et al. (1997) sont des exemples d’études qui
ont retenu une telle modélisation pluraliste, utilisant toutes deux d’une part une réponse
fonctionnelle du type Lotka-Volterra (comme exemple de contrôle de haut en bas) et
d’autre part une réponse fonctionnelle ratio-dépendante linéaire (pour représenter une
situation de contrôle par le donneur). La première étude était destinée à véri…er la
robustesse de certaines prédictions théoriques, tandis que la seconde a comparé les
46f
Identi…cation de modèles proie-prédateur (thèse de doctorat)
C. Jost
prédictions des deux modèles pour dé…nir des directions de recherche ultérieures et
interpréter des données de terrain.
Chapitre 3f
Conclusions générales et
perspectives
3.1f
Conclusions générales
Cette thèse peut être interprétée comme une validation générale d’un modèle proieprédateur ratio-dépendant. Les propriétés mathématiques de ce modèle ont été analysées
en détail. Cette analyse a montré que ce modèle peut prédire la coexistence stable de
la proie et du prédateur, la coexistence instable (cycle limite), l’extinction du prédateur seulement et l’extinction conjointe de la proie et du prédateur. Pour certaines
valeurs des paramètres, il existe deux bassins d’attraction, l’un pour l’origine (extinction) et l’autre pour une coexistence stable ou instable. Ceci explique naturellement
des résultats expérimentaux où l’on a observé que l’aboutissement à l’extinction ou à la
coexistence dépend des abondances initiales des abondances de proies et prédateurs. Il
est aussi démontré que le modèle ratio-dépendant explique correctement les dynamiques
qualitatives du plancton d’eau douce, des cultures de protozoaires et d’arthropodes en
conditions de laboratoire.
Dans une seconde étape, ce modèle ratio-dépendant est comparé qualitativement
et quantitativement à un modèle proie-dépendant de même complexité. La comparaison qualitative démontre que les deux modèles génèrent des patterns dynamiques très
similaires. La différence majeure est la réaction à un enrichissement, qui a un effet
déstabilisant et n’augmente que l’abondance à l’équilibre du prédateur pour le modèle
proie-dépendant, alors qu’il est neutre vis à vis de la stabilité et augmente les abondances
à l’équilibre de la proie et du prédateur pour le modèle ratio-dépendant. La comparaison
des deux modèles avec le modèle PEG (qui décrit verbalement les dynamiques observées
du plancton dans des lacs d’eaux douce de la zone tempérée) démontre qu’ils ne peuvent expliquer ces dynamiques que si l’on ajoute un effet saisonnier à un ou plusieurs
paramètres. Concernant les effets d’un enrichissement, il semble que le modèle ratiodépendant les explique légèrement mieux.
La méthode du maximum de vraisemblance s’est avérée être un fondement robuste pour la comparaison quantitative des deux modèles aux séries chronologiques
de systèmes proie-prédateur. Les données sur des protozoaires sont généralement mieux
47f
48f
Identi…cation de modèles proie-prédateur (thèse de doctorat)
C. Jost
décrites par le modèle proie-dépendant. Néanmoins, j’ai trouvé un système protozoaire
qui est signi…cativement mieux décrit par le modèle ratio-dépendant. Les données sur
des arthropodes sont mieux décrites par le modèle proie-dépendant. En revanche, les
deux modèles sont valides pour les données de phytoplancton-zooplancton. En conséquence, des prédictions pour des systèmes planctoniques devraient être basées sur les
deux modèles pour éviter d’être victime d’un artéfact mathématique de l’un d’entre eux
(par ex. le paradoxe de l’enrichissement).
En microbiologie, on trouve des formes équivalentes aux deux modèles, le modèle
de Monod (proie-dépendant) et le modèle de Contois (ratio-dépendant). Le type de
comparaison quantitative décrit ci-dessus a également été entrepris par des microbiologistes. En passant leurs résultats en revue, j’ai trouvé que le modèle de Monod semble
être correct avec des cultures en environnement homogène, des proies monospéci…ques
et des prédateurs monospéci…ques. Mais toute déviation par rapport à ces conditions
entraîne une prédateur-dépendance dans la fonction de croissance (réponse numérique),
de sorte que le modèle de Contois (un cas particulier de prédateur-dépendence) décrit
souvent mieux les dynamiques observées que le modèle de Monod. Le champ d’application du modèle de Contois se trouve aujourd’hui majoritairement dans des cultures
plurispéci…ques en milieu hétérogène, telles que celles des processus de fermentation ou
du traitement des eaux usés.
En conclusion, le modèle proie-dépendant semble approprié dans des cas de proie
monospéci…que et prédateur monospéci…que en milieu homogène. Dans toutes les autres
situations, divers degrés de prédateur-dépendance peuvent être observés et le modèle
faisant la meilleure approximation est a priori inconnu. Les concepts proie-dépendant
et ratio-dépendant peuvent être interprétés comme deux extrêmes par rapport à la
prédateur-dépendance. Le fait d’utiliser les deux modèles aide à la détection de prédictions qui sont sensibles à la prédateur-dépendance et à diriger des recherches approfondies si nécessaire.
3.2f
Perspectives pour la continuation des études
Les paragraphes suivants vont indiquer quelques possibilités sur la façon dont les travaux
présentés pourraient être continués.
Une approche non paramétrique
La technique employée pour la sélection de modèles dans les chapitres 7 et 8 demande
d’estimer tous les paramètres du modèle proie-prédateur, en particulier ceux de la fonction de croissance de la proie et le taux de mortalité du prédateur. Le but étant d’identi…er la forme de la réponse fonctionnelle, l’estimation des autres paramètres n’est
pas seulement sans interêt, mais peut même in‡uencer le résultat de la détection de la
prédateur-dépendance si les formes paramétriques de la croissance et de la mortalité
sont mal choisies. Un autre problème est l’absence d’effets retardés dans notre modèle
proie-prédateur. Un délai dans la réponse numérique par rapport à la réponse fonctionnelle est théoriquement plausible, et il a amélioré l’ajustement dans l’analyse de
C. Jost
Identi…cation de modèles proie-prédateur (thèse de doctorat)
49f
séries chronologiques de protozoaires (provenant des expériences classiques de Luckinbill 1973) effectuée par Harrison (1995) et la description qualitative d’un système
bactéries-bactériophage de Bohannan & Lenski (1997).
Une solution élégante à ces deux problèmes a été proposé par Ellner et al. (1997).
Les auteurs ont étudié le système
dx(t)
= f (x (t
dt
)) + g(x(t)).
Sur la base d’une approche de séries chronologiques, ils développent une méthode pour
estimer le retard et pour reconstruire f et g d’une manière non paramétrique. En
testant cette approche avec des données arti…cielles (auxquelles ils ajoutent une erreur
de processus et d’observation), ils trouvent qu’une estimation correcte de est possible
pour des séries chronologiques avec un minimum de 100 éléments. La reconstruction non
paramétrique de f et de g semble par ailleurs raisonnable, et les auteurs mentionnent
que leur méthode peut révéler des informations sur la biologie de f et g.
Ceci est exactement ce qui pourrait être fait pour identi…er la réponse fonctionnelle.
En estimant le retard dans la réponse numérique, on reconstruit aussi la réponse fonctionnelle de cette manière non-paramétrique avec des intervalles de con…ance pour toutes
les variables. Ensuite, on peut directement ajuster différentes formes paramétriques de
la réponse fonctionnelle à ces données reconstruites (en se servant de la méthode des
erreurs dans les variables, voir 3.2.3f) et identi…er empiriquement le modèle le plus
parcimonieux (utilisation de critères d’information). Des systèmes protozoaires tels que
ceux étudiés par Veilleux (1979) pourraient fournir les données nécessaires pour une
telle analyse.
3.2.1f
Modéliser l’extinction comme un phénomène déterministe
Le calcul du ‘quasi’-risque d’extinction d’une population naturelle se fait habituellement en paramétrisant un modèle de cette population (par exemple, avec un modèle
de Leslie 1948), en faisant des simulations stochastiques et en calculant la proportion
de simulations durant lesquelles la population est tombée en dessous d’un certain seuil
dé…ni au préalable. Une des raisons pour cette approche est que les modèles classiques
ne permettent pas l’extinction d’une manière déterministe. Le fait de travailler avec des
modèles incluant ce comportement (par ex. des réseaux trophiques ratio-dépendants
ou des réseaux trophiques basés sur les modèles proie-prédateur élaborés à la section
2.3.1f) pourrait fournir des informations supplémentaires concernant les risques d’extinction de populations menacées. L’extinction déterministe a été très peu considérée
dans la littérature écologique.
3.2.2f
Modélisation plus réaliste du plancton
Les données planctoniques du Lac Léman n’ont été analysées que dans le contexte proieprédateur, avec des facteurs exogènes constants. En fait, on dispose de beaucoup plus de
50f
Identi…cation de modèles proie-prédateur (thèse de doctorat)
C. Jost
données : les abondances de chaque espèce, la température à plusieurs profondeurs, le
contenu de l’eau en azote et en phosphore. En utilisant ces données comme information
supplémentaire, une prédiction une ou deux semaines à l’avance semble tout à fait faisable. Cependant, le fait d’ajouter des entrées va toujours améliorer l’ajustement, mais
aussi augmenter la complexité du modèle (avec tous les inconvénients que cela implique :
plus de paramètres à estimer, plus de difficultés pour transférer le modèle vers un autre
lac, etc.). De plus, l’amélioration de l’ajustement n’est pas nécessairement signi…cative
(‘critère d’information’, test du rapport des vraisemblances). L’identi…cation des variables externes les plus importantes est donc un vieux problème en écologie. L’approche
statistique habituelle utilise des techniques d’ACP. Néanmoins, ces techniques ne détectent que des relations linéaires. Les non-linéarités peuvent apparaître comme du
bruit. De méthodes d’ajustement empiriques telles que les réseaux neuronaux peuvent
aussi détecter des patterns non-linéaires (Lek et al. 1996). En divisant les données en
deux parties (‘in-sample’ et ‘out-of-sample’), on peut éduquer le réseau neuronal avec
la première partie puis tester son pouvoir prédictif avec l’autre partie. Cette procédure
peut être combinée avec les méthodes du bootstrapping. Avec ces méthodes, il semble
faisable d’identi…er les deux, trois ou quatres variables exogènes les plus importantes,
y compris par rapport aux dépendances non-linéaires. On pourra ensuite construire un
modèle dynamique simple pour prédire les variables intéressantes, par ex. le phyto- ou
zooplancton total.
3.2.3f
Comment tenir compte à la fois d’une erreur de processus et d’une erreur d’observation?
Toutes les méthodes statistiques utilisées dans cette thèse se basent sur l’hypothèse qu’il
y a seulement un type de stochasticité dans les données, soit une erreur d’observation,
soit un processus stochastique. Or, les données écologiques contiennent toujours les
deux stochasticités à la fois. Plusieurs stratégies possibles ont été développées par les
statisticiens pour contrer ce problème. Dans cette section je vais discuter trois stratégies
potentielles. Toutes ces stratégies sont numériquement plus coûteuses que les méthodes
décrites auparavant, mais avec la prochaine génération de microordinateurs le temps de
calcul va probablement diminuer pour atteindre un niveau acceptable.
Hilborn & Walters (1992) proposent d’ajuster les données comme si toutes les stochasticités provenaient de la même source (par ex., erreur d’observation ou de processus),
et d’évaluer les biais potentiels par simulation stochastique. Dans un cadre d’équations
différentielles, cela signi…e que l’on ajuste le modèle formulé en équations différentielles
ordinaires (EDO) à des séries chronologiques provenant d’une équation différentielle
stochastique ordinaire (EDSO). Cette approche semble parfaitement valable pour l’estimation de paramètres si aucune doute ne subsiste quant à la forme paramétrique du
modèle. En revanche, nous ne savons pas si l’espérance d’une EDSO garde la structure de son noyau déterministe sous la forme d’une EDO (ce qui est important pour
l’identi…cation correcte d’un modèle). Ceci est le cas pour les modèles de croissance
malthusienne (Yodzis 1989) et logistique (Nisbet & Gurney 1982; Braumann 1983),
mais ces résultats s’appuient sur l’existence d’une solution analytique de l’EDO. Je ne
connais pas de travaux de ce genre pour des modèles plus complexes, tels que les mod-
C. Jost
Identi…cation de modèles proie-prédateur (thèse de doctorat)
51f
èles proie-prédateur. Quelques études préliminaires indiquent qu’il peut y avoir une
différence entre modèle déterministe et l’espérance de l’EDSO : j’ai simulé des EDSO en
ajoutant un bruit blanc avec un écart-type proportionnel à l’abondance effective (simulation d’une erreur à caractéristiques log-normales, voir aussi le chapitre 7), en calculant
l’espérance à partir de beaucoup de réplicats (voir …gure 3f.1, losanges et étoiles). Le
fait d’ajuster l’EDO qui était le noyau de l’EDSO à cette espérance a démontré que,
souvent, les espérances ne concordaient plus avec le modèle déterministe, de sorte que
l’identi…cation de modèles devient plus difficile. La Figure 3f.1 indique des exemples
modérés de ces déviations du comportement déterministe, avec le meilleur ajustement
à l’EDO et la trajectoire de l’EDO avec les paramètres d’origine. Ces deux exemples
illustrent amplement que la structure de l’espérance d’une EDSO peut être différente
de celle de son noyau déterministe. Cependant, davantage de travaux numériques sont
nécessaires pour mieux déterminer de quelle manière un bruit stochastique peut changer
la structure d’une EDSO.
prey-dependent data
ratio-dependent data
(a)
abundances
(b)
time
time
Figure 3f.1: Deux exemples illustrant le problème que les espérances d’une EDSO
peuvent avoir une structure différente de celle du noyau déterministe (EDO) de
de l’EDSO. Les losanges et les étoiles représentent respectivement les espérances
des abondances de proies et de prédateurs (obtenus par simulation Monte Carlo),
la ligne continue est la trajectoire déterministe avec les paramètres originaux et
la ligne pointillée est le meilleur ajustement possible du modèle déterministe à
la trajectoire espérée. (a) est un exemple pour le modèle proie-dépendant, où
l’espérance est fortement différente de la trajectoire d’origine tandis qu’elle peut
encore être ajustée raisonnablement au modèle déterministe. (b) montre le cas d’un
modèle ratio-dépendent où l’espérance est plus proche de la trajectoire originale,
mais où il est difficile d’ajuster le modèle déterministe aux espérances.
La méthode des erreurs dans les variables (‘errors-in-variables’, le fait d’avoir une
incertitude dans les variables dépendantes et dans les variables indépendantes) offre
une autre approche au problème de l’ajustement de modèles dynamiques à des séries
chronologiques. L’idée est d’estimer non seulement les paramètres, mais aussi toutes les
variables indépendantes (‘nuisance parameters’). Ceci peut être fait dans le cadre d’une
approche de maximum de vraisemblance, mais la connaissance de l’une des deux erreurs
(observation ou processus) est indispensable. Ces méthodes ont été décrites pour des
modèles linéaires ou non-linéaires dans l’article fondamental de Reilly & Patino-Leal
52f
Identi…cation de modèles proie-prédateur (thèse de doctorat)
C. Jost
(1981). Les auteurs proposent des itérations emboîtées pour estimer alternativement
les paramètres et les variables d’état. Schnute (1994) et Schnute & Richards (1995)
discutent l’application de cette méthode dans le contexte de la gestion des populations
de poissons. Cette métode ne semble pas encore être utilisée dans d’autre branches de
l’écologie.
Une approche encore différente est constituée par les techniques de régression sur
la base de simulations (Ellner & Turchin 1995; Gouriéroux & Monfort 1996). L’idée
est d’utiliser un modèle stochastique pour simuler un jeu de données de référence sur
la base duquel on estime l’écart entre les vraies données et ce modèle (ces techniques
sont inspirées des techniques de reconstruction utilisées en théorie du chaos, voir Kantz
& Schreiber (1997) pour une introduction très lisible). Les paramètres dans le modèle stochastique sont sélectionnés de sorte que l’écart devienne minimal. Comme dans
la méthode des erreurs-dans-les-variables, cette technique demande la connaissance indépendante de l’un des deux types d’erreur. On en sait très peu sur la …abilité des
paramètres estimés et sur la sélection de modèle par cette technique lorsqu’elle est
utilisée avec des données écologiques typiques.
Part II
Detailed studies (accepted or
submitted articles)
49
Chapter 4
The clear water phase in lakes: a
non-equilibrium application of
alternative
phytoplankton-zooplankton models
Roger Arditi, Christian Jost, and Vojtěch Vyhnálek
51
52
Identifying predator-prey models (PhD Thesis)
C. Jost
Abstract
The verbal PEG-model (that describes the plankton dynamics of freshwater lakes of the
temperate zone) is interpreted in the context of dynamic predator-prey (phytoplanktonzooplankton) interactions. We then compare two explicit predator-prey models qualitatively with these dynamics. One model represents the ideas of top-down control
while the other model includes also ideas of bottom-up control. It is shown that neither model can explain the dyamics of the PEG-model satisfyingly, but that the second
model (bottom-up) has more in common with the observed dynamics than the …rst
model (top-down). We then study some extensions of both models (seasonality of parameters or additional interacting trophic levels) and show that both extended models
can predict the PEG-model dynamics satisfyingly. We conclude that the PEG-model
dynamics have, on the whole, more in common with the model containing bottom-up
control, but that we have to add a trophic level to the predator-prey framework (either as a seasonally changing parameter or as an additional state variable) to obtain
qualitatively satisfying dynamics.
Nous interprétons le modèle (verbal) “PEG”, qui décrit la dynamique du plancton
des lacs d’eau douce de la zone tempérée, dans le cadre d’interactions dynamiques proieprédateur (phytoplancton-zooplancton). Nous comparons ensuite qualitativement cette
dynamique avec deux modèles explicites de systèmes proie-prédateur. L’un d’entre eux
est fondé sur une idée de contrôle “descendant” des abondances au sein des niveaux
trophiques, tandis que le second inclut également une régulation “ascendante” des effectifs. Nous montrons qu’aucun des deux modèles ne peut rendre compte de façon satisfaisante de la dynamique décrite par le modèle PEG, mais que le second modèle (avec
régulation ascendante) présente plus de points communs avec elle que le premier (avec
contrôle descendant). Nous étudions ensuite un certain nombre d’extensions des deux
modèles (modi…cations saisonnières des valeurs des paramètres ou niveau trophiques
supplémentaires), et nous montrons que les deux modèles ainsi étendus peuvent prédire
la dynamique du modèle PEG de façon satisfaisante. Nous concluons que la dynamique
décrite par le modèle PEG a, dans l’ensemble, plus de points communs avec le modèle
incluant une régulation ascendante, mais qu’il est nécessaire de prendre en compte un
niveau trophique supplémentaire par rapport au cadre proie-prédateur (soit sous forme
d’un paramètre dont la valeur change de manière saisonnière, soit sous la forme d’une
variable d’état supplémentaire) a…n d’obtenir une dynamique satisfaisante du point de
vue qualitatif.
Arditi, Jost, Vyhnálek
4.1
Alternative phyto-zooplankton models
53
Introduction
It is becoming increasingly accepted that ecological processes in lakes are largely determined by trophic interactions (e.g., reviews in Carpenter 1988; Carpenter & Kitchell
1994). Therefore, plankton populations in lakes (or in simpler experimental systems)
have become ideal systems for testing the hypotheses and predictions of predator-prey or
food web theories (e.g., McCauley & Murdoch (1987), Sarnelle (1992), Mittelbach et al.
(1988)). Arditi & Ginzburg (1989) have suggested that the functional response might
often be approximated by a function of the prey-to-predator ratio instead of just the
prey density (prey dependence) as in classical predator-prey theory (e.g., Rosenzweig
& MacArthur (1963)). In the ensuing debate, many arguments have centered on the
ability of alternative models to explain observed patterns in the equilibrium abundances
(or long-term averages) of the various trophic levels in lakes of different productivities
(Arditi et al. 1991a; Gatto 1993; Sarnelle 1994; Abrams & Roth 1994; Mazumder 1994;
Akçakaya et al. 1995; Lundberg & Fryxell 1995; McCarthy et al. 1995) or in ad-hoc
experimental setups (Arditi et al. 1991b; Arditi & Saïah 1992; Ruxton et al. 1992;
Holmgren et al. 1996). Properties of systems out of equilibrium have been much less
studied. Particularly noteworthy is the work of Carpenter & Kitchell (1994) and Carpenter et al. (1994). Using arti…cial data as well as plankton time series from two lakes
over seven years, these authors have shown that modest observation errors make it impossible to identify reliably the underlying model. Here, we take a different approach
to the study of dynamic properties of alternative predator-prey models. Rather than
trying to …t models to noisy data, we will examine the ability of the models to generate,
in a qualitative way, a standard non-equilibrium pattern of lakes: the clear water phase
that occurs in the spring in lakes of the temperate region. This pattern is commonly
thought to be caused by predator-prey interactions.
4.1.1
Seasonality of plankton in lakes
In the last decades, limnologists have come to consider that the sequence of planktonic
events in lakes of the temperate zone exhibits a regular pattern that can be predicted to a
certain extent. The current knowledge about plankton seasonality is summarized in a set
of statements by the Plankton Ecology Group (the so-called PEG-model, Sommer et al.
1986). Three distinct periods can be distinguished during the growing season according
to changes in the phytoplankton biomass: (1) the spring peak, (2) the depression of
biomass (known as the clear water phase), and (3) the summer peak (see Figure 4.1a).
This pattern is more pronounced in meso- and eutrophic lakes than in oligotrophic ones.
The increase of phytoplankton biomass in the spring is caused by a fast growth of
algal populations due to a high concentration of nutrients and increasing light (Sommer
et al. 1986). The collapse of the phytoplankton bloom in the spring and the induction of
the clear water phase is interpreted as being caused by overgrazing of the phytoplankton
by the zooplankton (Lampert 1985; Lampert et al. 1986; Sommer et al. 1986). However,
nutrient limitation and subsequent sedimentation of algal cells were found to be the most
important factors of the phytoplankton collapse after the spring phytoplankton bloom in
several lakes: Lake Geneva, France-Switzerland (Gawler et al. 1988), Lake Constance,
54
Identifying predator-prey models (PhD Thesis)
C. Jost
biomass
phytoplankton
zooplankton
spring
summer
time
Figure 4.1: (a) Plankton dynamics as proposed by the PEG-model and (b) criteria
for the plankton dynamics of a model.
Germany-Switzerland-Austria (Weisse et al. 1990), Lake Søbygård, Denmark (Jeppesen
et al. 1990) and the Římov Reservoir, Czech Republic (Vyhnálek et al. 1991). In
addition, a lysis of algal cells can play a signi…cant role during this period, according to
observations from the Římov Reservoir (Vyhnálek et al. 1993). In those cases in which
the clear water phase is induced by nutrient limitation, zooplankton grazing still plays a
role in maintaining the biomass of phytoplankton at a low level. Nutrient concentration
increases due to the fast turnover through the zooplankton (Fott et al. 1980; Sommer
et al. 1986). Food limitation and …sh predation cause a decrease of zooplankton.
Therefore, conditions become favorable for a second bloom of phytoplankton, followed
by an increase of zooplankton in the summer. Finally, reduced light availability results
in a decline of phytoplankton followed by a decline of herbivores towards winter.
These quantitative variations are accompanied by a qualitative succession in the
nature of species. The spring peak of phytoplankton is formed mostly by small fastgrowing algae easily edible by herbivorous zooplankton. This is dominated by large
species (especially Cladocera). During the clear water phase, large, colonial, inedible,
and toxic phytoplankton species are favored because of their resistance against herbivore
grazing (Porter 1977). These species (especially cyanobacteria) become dominant and
form the summer peak of phytoplankton biomass. Large herbivores are replaced by
smaller species, less vulnerable to predation, and less affected by perturbation of their
feeding apparatus by large algae. Thus, the species composition of both trophic levels
changes through the seasons under the in‡uence of predation and resource competition
(Sommer et al. 1986).
Arditi, Jost, Vyhnálek
4.1.2
Alternative phyto-zooplankton models
55
Simple mathematical models of predator-prey interactions
The general model describing the dynamics of prey-predator populations in continuous
time can be written as:
dN
dt
dP
dt
= f (N )N
g(N, P )P
= eg (N, P )P
P
(4.1)
(4.2)
where N is prey abundance, P is predator abundance, t is the time, f (N ) is the per
capita net prey production in the absence of predation, g(N, P ) is the functional response of predators (the number of prey killed by one predator in a unit of time), e is
the conversion efficiency and is the per capita death rate of predators. The key role in
prey-predator models is played by the functional response g (Solomon 1949), sometimes
called the trophic function (Svirezhev & Logofet 1983). Traditionally, it is assumed
that the functional response g is a function of prey density only [prey-dependent feeding, g = g(N )], without any dependence on predator density (Holling 1959a; Rosenzweig
1971; May 1973). The hypothesis g = g(N ) is based on an analogy with the law of mass
action in chemistry assuming that prey and predator individuals encounter each other
randomly in space and time (Royama 1971). Therefore, the prey-dependent model can
be applied to systems which are spatially homogeneous and in which the time scale of
prey removal by predators is of the same order of magnitude as that of population reproduction (Arditi & Ginzburg 1989). These conditions are ful…lled especially in small-scale
and well-mixed laboratory systems containing bacteria (Jannasch 1967; Luckinbill 1973;
Bazin 1981), algae (Droop 1966; Goldman 1977), protozoa (Taub & McKenzie 1973),
and under certain conditions also pelagic rotifers (Droop & Scott 1978; Boraas 1980)
and cladocera (Arditi et al. 1991b; Arditi & Saïah 1992).
However, natural ecosystems are usually spatially heterogeneous and the time scales
for feeding and reproduction are also often very different. If the spatial heterogeneity
can be characterized by a double exponential distribution of the encounter time, Ruxton
& Gurney (1994) showed that a purely prey-dependent functional response may still be
derived. However, various other mechanisms (e.g., pseudo-interference, etc.) might
lead to explicit dependence on predator density [g = g(N, P )]. Arditi & Ginzburg
(1989) have argued that, in many cases, this predator dependence could be simpli…ed
as a ratio-dependent model [g = g(N/P )] instead of modeling explicitly all conceivable
interference mechanisms (and thus adding parameters to the model). A most striking
difference between prey-dependent and ratio-dependent models is the response of the
equilibria of trophic levels after an increase in prey production. The prey-dependent
model predicts an increase of predator abundance only, with prey abundance remaining
unchanged. In systems consisting of more than two trophic levels, the prey dependent
model predicts various responses of the several levels to an increase of primary input,
depending on the chain length and on the level considered: no response, proportional
response, non-linear increasing response and non-linear decreasing response (Oksanen
et al. 1981; Persson et al. 1988). On the other hand, the ratio-dependent model predicts
proportional responses of all trophic levels to an increase of primary input (Arditi &
56
Identifying predator-prey models (PhD Thesis)
C. Jost
Ginzburg 1989).
Recent analysis of data from freshwater ecosystems have shown evidence that average biomasses of …sh, zooplankton and phytoplankton are positively correlated along a
gradient of productivity (summarized by Arditi, Ginzburg, & Akçakaya 1991a; Ginzburg
& Akçakaya 1992; Mazumder 1994). These positive correlations between trophic levels
are in agreement with the predictions of the ratio-dependent model if the equilibria of
all trophic levels are approximated by average biomasses over the whole year or over
the growing season. On the opposite, the empirical …ndings from freshwater ecosystems
are in contradiction with the prey-dependent model predicting a mixture of positive,
negative and zero correlations between trophic levels (Oksanen et al. 1981; Arditi &
Ginzburg 1989; Arditi, Ginzburg, & Akçakaya 1991a), and also a density-dependent
mortality rate as proposed by Gatto (1991) predicts these positive correlations only for
two-level systems, not for food chains (see 4.B).
Further insights into this controversy may be gained by comparing directly the trajectories of the two models with time series from real ecosystems. This may be done
by …tting the models to the data and using goodness-of-…t methods as a criterion.
Carpenter et al. (1994) found that by this method the two models could be reliably
distinguished only if the ecosystems underwent considerable perturbations (e.g. caused
by introducing or removing …sh populations). Applying this method to such a dataset
[zooplankton and edible phytoplankton in Tuesday lake (Carpenter & Kitchell 1994)]
they detected a slightly better …t of the ratio-dependent model.
Another method to compare the trajectories is the mathematical analysis of the dynamics. Despite numerous results on the equilibrium properties of the two alternative
models, our knowledge about the non-equilibrium properties is limited. In this paper
we present a comparison of the transient behavior of both models before reaching the
equilibrium. Then, the trajectories of population densities predicted by the models
are compared with the seasonal development of the plankton community in lakes of
the temperate zone, which is understood as a process reaching equilibrium (or quasiequilibrium) during the summer. Furthermore we study the qualitative changes of the
system dynamics if the model parameters change (during season or with eutrophication). Suggested seasonal changes are decreasing predator attack rate (due to increased
proportions of inedible algae), increasing predator mortality rate (due to predation by
higher trophic levels) and increasing prey growth rate (due to higher light intensity and
water temperature). The effects of eutrophication are studied by increasing the carrying
capacity (see McCauley et al. 1988).
4.2
4.2.1
Problem formulation
The alternative models
Two-level models consisting of phytoplankton as prey (N ) and zooplankton as predators (P ) will be built with differential equations of type (4.1–4.2). The production of
Arditi, Jost, Vyhnálek
Alternative phyto-zooplankton models
57
phytoplankton follows the usual logistic growth:
f (N ) = r(1
N
).
K
(4.3)
The functional response of zooplankton is described by a concave monotonic upperbounded expression. In the prey-dependent model, it is a function of prey density N
(type II, Holling 1959a):
g(N, P ) = g(N ) =
aN
,
1 + ahN
(4.4)
where a is the searching efficiency and h is the handling time.
In the ratio-dependent model, g is a function of the ratio N/P :
g(N, P ) = g
N
P
=
N/P
N
=
,
1 + hN/P
P + hN
(4.5)
where has different units from a.
4.2.2
The required patterns
In order to be satis…ed that a simple prey-predator model approximates the complex
succession of events of the PEG-model (Sommer et al. 1986), the following conditions
are required. These criteria are set mainly for the dynamics of phytoplankton but
zooplankton dynamics can be used as a secondary criterion.
1. Starting with a low (winter) biomass, the phytoplankton must present a (spring)
maximum followed by a distinct depression (the clear water phase) and then a
high biomass (during the summer) (Fig. 4.1).
2. A relatively constant biomass of phytoplankton is expected thereafter.
3. Zooplankton dynamics, starting from a low biomass, must follow the spring phytoplankton increase with some delay and a stabilization of zooplankton biomass
is expected in the summer.
4. In lakes with a higher degree of eutrophication, i.e., with a higher prey carrying
capacity, the model should show wider amplitudes and a higher algal equilibrium.
Thus, a good qualitative similarity between the simulated dynamics and the PEGmodel is required during the spring and the summer. In the autumn, both phytoplankton and zooplankton are expected to remain at steady levels in the simulations because
the decline that occurs in nature is assumed to be caused by external physical factors
(decreasing light and temperature).
Identifying predator-prey models (PhD Thesis)
58
4.3
4.3.1
C. Jost
Model analysis and results
Dimensionless forms
With appropriate changes in variables, the two models built on the functional responses
(4.4–4.5) can be simpli…ed to dimensionless forms. This reduces the number of independent parameters and makes the mathematical analysis easier. The prey-dependent
model becomes
dNp
dt0
dPp
dt0
= R(1
=
Np
)Np
C
Np
Pp
1 + Np
Np
Pp
1 + Np
QPp ,
(4.6)
(4.7)
with Np = ahN , Pp = ahP/e, t0 = et/h, R = rh/e, C = ahK and Q = h/e.
The dimensionless form of the ratio-dependent model is
dNr
dt0
dPr
dt0
Nr
Nr
Pr
)Nr D
D
Pr + DNr
Nr
Pr QPr ,
= D
Pr + DNr
= R(1
(4.8)
(4.9)
where Nr = hN/(eK ), Pr = hP/(e2 K ) and D = h/e.
4.3.2
Isoclines
The predator isocline is a straight line, which is vertical in the prey-dependent model
and slanted (through the origin) in the ratio-dependent model (Fig. 4.2). The prey
isocline is always parabolic in the prey-dependent case. Its “hump” may be in the
positive quadrant or not. However, the latter case is considered atypical (Rosenzweig
1969). In the ratio-dependent model, the prey isocline is usually parabolic-like but it
may also have a vertical asymptote [the “limited predation” case in Arditi & Ginzburg
(1989)]. In either case, it will always be entirely in the positive quadrant.
Ratio dependence
predator
predator
Prey dependence
prey
prey
Figure 4.2: Typical isoclines of the prey-dependent model (left) and the ratiodependent model (right). The humped line is the prey isocline, the straight line is
the predator isocline.
Arditi, Jost, Vyhnálek
4.3.3
Alternative phyto-zooplankton models
59
Equilibria
Besides the trivial equilibria (0, 0) and (0, 1), both models have one non-trivial equilibrium which is, in the prey-dependent model,
Np =
Pp
Q
(4.10)
1 Q
R =
N (C
CQ p
Np ) =
R(C Q CQ)
.
C(1 Q)2
(4.11)
and in the ratio-dependent model
D
(DQ + R D)
R
R D2
Nr (D Nr ) =
=
(2DQ
DQ
QR
Nr =
Pr
(4.12)
DQ2 + R
QR
D).
(4.13)
Conditions for the positiveness of the equilibrium are thus
0<
Q
1
Q
<C
(4.14)
R
D
(4.15)
in the prey-dependent model and
0<1
Q<
in the ratio-dependent model.
In the prey-dependent model, when the parameter C approaches the bound Q/(1
Q), the prey equilibrium remains at Q/(1 Q) while the predator equilibrium tends to
0. In the ratio-dependent model both populations go extinct when 1 Q approaches the
upper bound R/D, while the prey equilibrium tends to D and the predator equilibrium
to 0 for Q approaching 1.
4.3.4
Stability of the non-trivial equilibrium
In both models, the non-trivial equilibrium point can be locally stable or unstable. With
(mij ) being the community matrix (the Jacobian at the equilibrium point, see 4.A for
further details), the characteristic equation for the eigenvalues reduces to a quadratic
equation in the case of a two-level model:
2
(m11 + m22 ) + m11 m22
m12 m21 = 0.
(4.16)
The Routh-Hurwitz criterion (e.g., Amann 1990) says that the equilibrium point is
stable whenever the two conditions
m11 m22
m12 m21 > 0
m11 + m22 < 0
(4.17)
(4.18)
Identifying predator-prey models (PhD Thesis)
60
C. Jost
hold simultaneously (see 4.A for further details).
In the case of the prey-dependent model, the stability conditions are
Q
1
Q
<C<
1+Q
1 Q
(4.19)
where the lower bound coincides with the condition for existence of the positive equilibrium. In the ratio-dependent model, we have a stable positive equilibrium point
whenever the conditions
D<
Q
Q2 + R
1 Q2
and
1
Q<
R
D
(4.20)
are ful…lled.
The stable equilibrium may be reached with or without oscillations in both models.
Oscillations occur if the discriminant of equation (4.16) is negative, i.e., if
(m11
m22 )2 + 4m12 m21 < 0.
(4.21)
This criterion gives rather complicated upper and lower bounds for C (respectively D)
which the interested reader may …nd in 4.A. More important than the formulas is the
visualization in parameter portraits (Fig. 4.3, see next subsection).
4.3.5
Comparison of dynamic properties in the two models
The criteria for positive equilibrium point, stability (attractor or repellor) and oscillations (spiral or non-spiral) can be expressed in a graphic form as parametric portraits
of the parameters C (resp. D) and Q after …xing R as a constant. In order to identify
the effect of changing the growth rate r, another parametric portrait has been created
where C (resp. D) is plotted versus R while keeping Q constant (see Fig. 4.3).
Several interesting properties of the models can be pointed out:
The existence of positive equilibria requires in both models that Q < 1 holds or,
in the original parameters, that h < e.
There is a fundamental difference regarding the size of the searching efficiency
a (resp. ): while in the case of prey-dependence, it has a lower bound (a >
/[K (e h)] because C > Q/(1 Q)), in the case of ratio-dependence it has an
upper bound ( < re/(e h) because D < R/(1 Q)).
For C 1 (resp. D 1), in both models the positive equilibrium is stable.
However, for C > 1 (resp. D > 1 ), the equilibrium can be either stable or unstable.
In the initial parameters, this means that an unstable positive equilibrium can exist
in the prey-dependent model for ahK > 1 and in the ratio-dependent model for
h > e.
Arditi, Jost, Vyhnálek
Alternative phyto-zooplankton models
Prey dependence
C=a b K
5
61
Ratio dependence
D=a b/e
5
1
4
4
3 2 1
Q=b mu/e
4/(4+R)
3
2
Q=b mu/e
1
Prey dependence
C=a b K
5
4
1
1-R
1
Ratio dependence
D=a b/e
5
1
4
3
2
3
1
R=r b/e
4(1-Q)/Q
2
R=r b/e
1-Q
Figure 4.3: Parametric portraits of C versus Q for …xed R (upper left) and of C
versus R for …xed Q (lower left) for the prey-dependent model and of D versus
Q for …xed R (upper right) and of D versus R for …xed Q (lower right) for the
ratio-dependent model. In regions 1 populations go extinct, in regions 2 the stable
equilibrium point is reached without oscillations, in regions 3 the equilibrium point
is a spiral node, in regions 4 the positive equilibrium point is a spiral repellor and
in regions 5 the positive equilibrium is an ordinary repellor.
In the ratio-dependent model, unstable equilibrium points are possible only for
D > R, while for D R only stable equilibrium points exist. The reason is
that if r (⇔ D R), the prey isocline has a vertical asymptote and the
equilibrium point is always stable (Arditi & Ginzburg 1989). On the other hand,
if > r (⇔ D > R), the prey isocline is “humped" and either stable or unstable
equilibrium points can be found. Note that equilibria can be stable even if they
lie on the ascending part of the hump in contrast to an erroneous assertion in
Arditi & Ginzburg (1989) (see Jost et al. (1999) for an analytical proof). No
similar condition is present in the prey-dependent model, which has only humped
isoclines, and equilibria lying on the ascending part of the prey isocline are always
unstable.
In the prey-dependent model, since C = ahK , increasing the carrying capacity K
in Eq. (4.3) acts as a destabilizing factor (see Fig. 4.3: increasing K increases C ,
thus leaving the area of stable positive equilibrium). This fact is well known as the
“paradox of enrichment” (Rosenzweig 1971) (see Fig. 4.4a). No such destabilizing
effect of K exists in the ratio-dependent model: K is used (in the dimensionless
form) as a scaling factor only for Nr and Pr , but not for any of the parameters
D, Q, or R. Therefore, the only in‡uence of increasing K is to increase both
62
Identifying predator-prey models (PhD Thesis)
C. Jost
populations N and P in the same proportion (see Fig. 4.4b).
The prey growth rate r does not in‡uence stability in the prey-dependent model,
since changing r moves the parameter pair (C, R) along a horizontal line (lower
graphs in Fig. 4.3) which does not intersect any of the stability bounds (it may
only in‡uence the oscillatory behavior). In the ratio-dependent model, on the
contrary, for any Q < 1 and any D, increasing r will stabilize the equilibrium,
while decreasing r will make the system leave the area of positive equilibrium
points (extinction of both populations).
Prey dependence
Ratio dependence
prey
prey
(a)
time
(b)
time
Figure 4.4: In‡uence of K on the stability, shown here for the prey dynamics
(higher amplitudes with higher K ), for the prey-dependent model (a) and for the
ratio-dependent model (b).
4.3.6
Model trajectories and plankton seasonality
According to the criteria 1–2 stated above (Fig. 4.1), plankton dynamics must reach
a stable equilibrium after one oscillation. Therefore, the trajectories obtained in the
areas of stable equilibrium points reached after oscillations (regions 3 in Fig. 4.3) must
be examined. Typical prey (phytoplankton) trajectories in this region are presented in
Fig. 4.5 for both models. A distinct maximum is formed if the starting biomass is low.
Then, a deep minimum can follow. However, in such case, a low equilibrium point is
reached after many oscillations, gradually decreasing in amplitude. If it is attempted
to reach a high steady state after few oscillations, this can only be the case with a very
mild depression of the prey biomass after the …rst maximum (see Fig. 4.5). That there
can only be a gradual decrease of amplitudes and not a …rst strong oscillation followed
by small oscillations around a high equilibrium is intuitively apparent for both models
already when looking at the isocline portraits (Fig. 4.2).
The predator (zooplankton) exhibit similar trajectories, lagging behind the prey
trajectory. In this respect criterion 3 can be considered satis…ed by both models.
With respect to criteria 1–2, it is clear that neither the prey-dependent nor the
ratio-dependent model are able to generate satisfactory trajectories. Both are unable to
generate the spring clear water phase, along with a stable equilibrium in the summer,
of a magnitude similar to the …rst peak and reached after few oscillations.
Arditi, Jost, Vyhnálek
Alternative phyto-zooplankton models
Prey dependence
63
prey
prey
Ratio dependence
(a)
(b)
time
time
Figure 4.5: Trajectories of prey in the area of stable equilibrium points reached
with oscillations (prey-dependent model (a) and ratio-dependent model (b)). The
highly oscillating trajectory was produced with parameters near the criterion of
stability, the other with parameters near the criterion of oscillations.
However, increasing eutrophication (increasing K ) behaves differently in both models (Fig. 4.4). Increasing K has a destabilizing effect (higher amplitudes, increasing
frequency, …nally sustained oscillations) with an unchanged prey equilibrium in the
prey-dependent model. In the ratio-dependent model, there are no effects on stability
properties; only a quantitative, proportional increase in the amplitudes and in the level
of the prey equilibrium. With respect to criterion 4, the ratio-dependent model behaves
therefore much better than the prey-dependent model.
4.3.7
Seasonal changes of parameter values
The above analysis assumed autonomous equations, i.e., time-independent parameters.
However, it can be argued that one or several parameters must change along the seasons. First, it is most reasonable to assume that the phytoplankton intrinsic growth
rate r increases as temperature increases from spring to summer. Second, zooplankton
parameters may also change: the attack rate a (resp. ) can decrease as a result of an
increased proportion of inedible algae and the mortality rate can increase because of
increased …sh predation. Using the expressions for the non-trivial equilibria (4.10–4.13),
and inspecting Fig. 4.3, it can be seen that variations of the parameters as just suggested have the effects summarized in Table 4.1. Variations in a (resp. ) or in have
in both models more or less the same effect, i.e., an increase of the prey equilibrium and
a stabilization, which is in agreement with our required criteria. Increasing the prey
growth rate r has practically no qualitative effect in the prey-dependent model, while
it has a desirable effect in the ratio-dependent model (i.e., an increase of the prey equilibrium and a stabilization). Given these properties, it is not difficult to …nd scenarios
with varying parameters that will generate the desired trajectories with either model.
Examples are given in Fig. 4.6.
64
Identifying predator-prey models (PhD Thesis)
C. Jost
Table 4.1: Summary of the possible trends of parameters during the season or with
eutrophication and their effects on system stability and equilibria.
effects in
Parameter Cause
Time
preyratiovaried
scale
dependence
equ. stab.
equ. stab.
2+
Eutrophication
Decades
K%
&1+
%
Temp. increase
Seasonal
r%
%
%
increase of
Seasonal
a&
%
%
%
%
inedible algae
( &)
higher predation Seasonal
%
%
%
%
%
1+ :
increasing frequency and amplitude
same frequency, increasing amplitude
: with further change the parameters leave the area of positive equilibrium
2+ :
4.4
Discussion
The most important difference between the prey-dependent and ratio-dependent twolevel models is the fact that an increase of prey carrying capacity K causes an increase
of both prey and predator equilibrium densities in the ratio-dependent model, whereas
only predator equilibrium density increases in the prey-dependent model (Arditi &
Ginzburg 1989). This property at steady-state conditions has to be considered also when
studying nonsteady-state characteristics of both models. Moreover, general plankton
dynamics seem to be relatively independent of trophic input, because similar seasonality
of plankton was found in lakes of different levels of eutrophication. This enabled the
formulation of the PEG-model (Sommer et al. 1986). This property is only found in
the ratio-dependent model, where changes of K , the main parameter changing with the
trophic input, have no in‡uence on the qualitative behavior, only on the amplitudes
(see …gure 4.4). On the other hand, the stability of the prey-dependent model depends
on K and the trajectories can be qualitatively different for different values of K .
Prey dependence
Ratio dependence
plankton
(b)
plankton
(a)
time
time
Figure 4.6: Resulting trajectories if the predator death rate is increasing during
the season. Both the prey-dependent (a) and the ratio-dependent (b) model may
thus show the desired trajectories.
Arditi, Jost, Vyhnálek
Alternative phyto-zooplankton models
65
However, despite the mentioned slight advantages of the ratio-dependent model,
both models give qualitatively similar trajectories which cannot ful…ll our criteria for
plankton seasonality. Their basic property is a gradual decrease of the amplitude of
oscillations after the …rst maximum, which is affected by the starting conditions. It is
impossible to obtain a distinct clear water phase period followed by a sudden decrease
of amplitude and fast reaching of stable equilibrium. The desirable trajectories in twolevel models can only be obtained if some parameter (or parameters) change during the
season, shifting the system from the area of unstable to the area of stable equilibrium.
An analogous approach was used in (McCauley et al. 1988) to explain the proportional
growth of phytoplankton and zooplankton with increasing productivity on the basis
of prey-dependent models. They assume the following trends with increasing nutrient
status of lakes: 1. the realized per capita growth rate of phytoplankton (both edible
and inedible) increases, 2. Daphnia’s attack rate decreases in response to increasing
concentration of inedible algae, 3. Daphnia’s death rate increases in response to greater
predation pressure.
Similar trends can be expected also during a growing season. A seasonal shift from
edible to inedible algae has been documented in many lakes (Sommer et al. 1986).
Colonial cyanobacteria often dominate in lakes exceeding a certain level of eutrophication during a summer period (Sas 1989). Therefore, a decrease of the attack rate of
Daphnia can be expected in the summer. An increase of the maximum growth rate of
phytoplankton (r) is probable due to a rise of temperature and light. The death rate
of herbivorous zooplankton is probably extremely variable and changing in the spring,
while in the summer it seems more uniform (Seda 1989). Lampert (1978) found in Lake
Constance, that the Daphnia population is not controlled by food in the spring but by
adult individuals of carnivorous Cyclops vicinus. The spring maximum of Daphnia can
develop only when Cyclops dies out. Feeding activity of planktivorous …sh increases
during the season along with increasing temperature. In addition, some species do not
feed during reproduction in the spring (...). Therefore, a general increase of grazing
pressure on herbivores during the season (thus an increasing predator mortality rate
in our models) can be expected, probably with dramatic short-term changes in the
spring.
Gatto (1991) proposes to introduce a density dependent mortality of predator (P ).
After this modi…cation, the prey-dependent two-level model predicts a proportional increase of both equilibrium densities, but the trajectories still exhibit qualitatively similar
properties as in the standard prey-dependent models, especially a gradual decrease of
the amplitude of oscillations after the …rst maximum. Moreover, when applying this to
three levels the positive correlations in general only hold for the top two levels, not for
the lower two levels under consideration (see 4.B).
One might suggest that the clear water phase is the result of cycling populations
rather than populations reaching equilibrium. Several arguments speak against this
interpretation. Sommer et al. (1986) analyzed 24 different lakes and interpreted the
high summer abundances as being in steady state. Furthermore, there are no reported
limit cycles of plankton in tropic lakes, which are permanently in the summer state of
lakes in temperate zones.
A further explanation for the insufficiency of a two-level model for application in
66
Identifying predator-prey models (PhD Thesis)
C. Jost
plankton dynamics can be that some processes not included in the model play an important role. It can be true especially in the spring, generally understood as a disequilibrium period (McCauley et al. 1988). Physical and chemical parameters change suddenly
and these changes can in‡uence the development of organisms in water. Vyhnálek et al.
(1994) found that the spring bloom of phytoplankton grown in the canyon-shaped Římov Reservoir is formed by phytoplankton grown in the head of the reservoir and then
drifted downstream to the main lake. This process accelerates spring dynamics of phytoplankton. The collapse of phytoplankton and the induction of the clear water phase
appears to be the key problem of plankton dynamics. In the PEG-model, it is interpreted as a consequence of prey-predator interactions (Sommer et al. 1986). However,
several authors found that the sudden crash of phytoplankton is a consequence of nutrient limitation and subsequent sedimentation of phytoplankton (Gawler et al. 1988;
Jeppesen et al. 1990; Weisse et al. 1990) or even of lysis of algal cells (Boekel et al.
1992; Vyhnálek et al. 1993). In these cases grazing of herbivores is a parameter of
second rate importance and the depression of phytoplankton biomass (the clear water
phase) is not caused by prey-predator relationships.
In summary, neither the prey-dependent nor the ratio-dependent prey-predator models composed of phytoplankton and zooplankton are able to predict seasonal dynamics
of these two levels. Nevertheless, the ratio-dependent model seems to be more realistic
in view of the following properties: 1. both phytoplankton and zooplankton biomasses
at equilibrium increase with increasing productivity and 2. there is no effect of productivity on the stability of the system. To obtain the expected prediction, with a
clear water phase and a stable equilibrium rapidly reached in the summer, the following
modi…cations of the two-level ratio-dependent model are proposed: 1. parameters of
phytoplankton and zooplankton change during the season, 2. introduction of a third
level (carnivorous zooplankton or …sh), 3. introduction of phytoplankton mortality other
than zooplankton grazing (sedimentation, lysis).
Acknowledgements
This research was supported by the Swiss National Science Foundation and by the
French ‘Programme Environnement, Vie et Société’ (CNRS).
Appendix
4.A
Detailed matrix analysis
In this chapter we will develop the detailed Jacobian matrices for both (dimensionless)
models. Since we will analyze one model after the other we will omit the subscripts p
for prey-dependent and r for ratio-dependent for easier notation.
Arditi, Jost, Vyhnálek
4.A.1
Alternative phyto-zooplankton models
67
Prey-dependent model
The differential equations are given by
dN
dt
dP
dt
N
)N
C
= R(1
=
N
P
1+N
N
P
1+N
(4.22)
(4.23)
QP.
This system has the two trivial equilibria (0, 0) and (C, 0) and the non-trivial equilibrium
Q (C Q CQ)R
(4.24)
.
,
1 Q
C (1 Q)2
The Jacobian is
"
R
2N R
C
P
(1+N )
P
(1+N )2
+
NP
(1+N )2
#
N
1+N
N
1+N
R
0
Q
(4.25)
.
0
, thus (0, 0) is a saddle
Q
At the equilibrium (0, 0) this matrix evaluates at
C
R
1+
C
point. At (C, 0) we get
, thus (C, 0) is stable whenever
C
0
Q
1+C
the non-trivial equilibrium point the Jacobian is
"
# Q(1 C +Q+CQ)R
Q
m
m
11
12
C(Q 1)
=
m21 m22
0
R QR QR
C
C
1+C
< Q. At
(4.26)
and its stability analysis is done in the text, while the oscillation criterion,
(m11
resolves to
m22 )2 + 4m12 m21 < 0,
(4.27)
p
2Q3 + QR Q3 R 2 (1 Q)3 Q(Q Q2 + R + QR)
< C < (4.28)
(1 Q)2 (4(Q 1) + QR)
p
2Q + 4Q2 2Q3 + QR Q3 R 2 (1 Q)3 Q(Q Q2 + R + QR)
. (4.29)
(1 Q)2 (4(Q 1) + QR)
2Q + 4Q2
The …rst expression tends to a …nite value for Q → 4+4R , while the second expression is
not de…ned at this value. Therefore the …rst inequality holds over the whole range of Q
(0 < Q < 1), while the second inequality only applies for Q between 4+4R and 1.
4.A.2
Ratio-dependent model
The differential equations are given by
dN
dt
dP
dt
N
)N
D
N
P
= D
P + DN
= R(1
D
N
P
P + DN
QP.
(4.30)
(4.31)
Identifying predator-prey models (PhD Thesis)
68
C. Jost
This system has the two trivial equilibria (0, 0) and (D, 0) and the non-trivial equilibrium
D(DQ + R D) D2 (1 Q)(DQ + R D)
,
.
(4.32)
R
QR
The Jacobian is
"
R
2N R
D
DP
D2 NP
+ (DN
+P )2
DN +P
DP 2
(DN +P )2
D2 N 2
(DN +P )2
D2 N 2
Q
(DN +P )2
#
.
(4.33)
1
R
At the equilibrium (D, 0) this matrix evaluates at
, thus (D, 0) is
0
1 Q
stable whenever 1 < Q (i.e. when the predator isocline has a negative slope and there
is no non-trivial positive equilibrium for prey and predator).
At the non-trivial equilibrium point the Jacobian is
D DQ2 R
Q2
D(1 Q)2
Q(Q 1)
(4.34)
and its stability analysis is done in the text, while the oscillation criterion,
(m11
m22 )2 + 4m12 m21 < 0,
(4.35)
resolves to
Q + 3Q2
3Q3 + Q4 + R
Q + 3Q2
p
Q2 R 2 (Q
(1 Q2 )2
3Q3 + Q4 + R
1)3 Q2 (Q
p
Q2 R + 2 (Q
(1 Q2 )2
Q2
1)3 Q2 (Q
QR)
R
Q2
R
< D < (4.36)
QR)
.(4.37)
The equilibrium (0, 0) is less easy to analyze since neither the differential equations
nor the Jacobian are de…ned at this point. Since the local behavior depends on how
the ‡ow looks like close to (0, 0) we reformulate our equations for the variables P and
L := N
:
P
dL
= L(R + Q
dt
dP
LP
= D
1 + DL
dt
R
LP )
D
QP.
D
L(1 + L)
1 + DL
(4.38)
(4.39)
This new system has three equilibria, a non-trivial one that shows the same behavior as
the non-trivial equilibrium in the original system, and two equilibria on the L axis, (0, 0)
and (0, DD(Q+QR R1) ). The former is stable whenever D > Q + R and the latter is a saddle
point whenever the non-trivial equilibrium is positive in both variables, as may be seen
on the isocline graph. Applied to the original system this means that for D < Q + R
and if there exists a positive non-trivial equilibrium (0, 0) is a saddle point, else it is
attractive for all trajectories where N/P approaches 0.
Arditi, Jost, Vyhnálek
4.B
69
Alternative phyto-zooplankton models
Some effects of a density dependent mortality
rate
Gatto (1991) and Gleeson (1994) pointed out that in food chains with Lotka-Volterra
functional responses and a toppredator with a mortality rate that is proportional to
its density the equilibrium densities of all trophic levels are correlated with primary
productivity. We will study the equilibrium behavior in a more general food chain
with three trophic levels, prey-dependent functional responses and a general densitydependent mortality rate of the top predator,
dP
dt
dH
dt
dC
dt
= f (P )P
g1 (P )H
(4.40)
= e1 g1 (P )H
g2 (H )C
(4.41)
= e2 g2 (H )C
(C )C,
(4.42)
where P are plants, H are herbivores and C are carnivores. f (P ), g1 (P ) and g2 (H ) are
increasing, bounded functions of their respective arguments. We assume that plant and
herbivore natural mortality is negligible compared to the mortality due to predation
and that the carnivore abundance in never close to 0.
In most models, the carnivore mortality rate is assumed to be a constant, (C ) = .
Here we assume that (C ) is some increasing, bounded function (due to passive or active
direct inhibition of competitors or if this trophic level is itself subject to predation by
some higher predator with Holling type III functional response (Holling 1959a)).
Setting (4.42) to 0 and solving for C we get
C = 1 (e2 g2 (H )),
(4.43)
thus at equilibrium C is positively correlated with P . Solving (4.41) for P we get
P = g1 1 (
g2 (H ) 1 (e2 g2 (H ))
.
e1 H
(4.44)
For small H g2 may be approximated by some linear function, g2 (H ) cH . In this case
we get
P = g1 1 (
c 1 (e2 cH )
,
e1
(4.45)
therefore P and H are positively correlated for small H . For large H g2 may be approximated by some constant, g2 (H ) c. This gives
P = g1 1 (
c 1 (e2 c)
),
e1 H
(4.46)
therefore P is negatively correlated with H . We may conclude that a density dependent
mortality rate of the top predator only leads to positive correlations between equilibrium
abundances of the top two levels, nothing speci…c can be said about correlations with
the lowest level.
70
Identifying predator-prey models (PhD Thesis)
C. Jost
Figure 4.7: Dynamics of total phytoplankton, herbivorous zooplankton and temperature in 1989. The developpement is representative for all years.
Chapter 5
About deterministic extinction in
ratio-dependent predator-prey
models
Christian Jost, Ovide Arino, Roger Arditi
(Bulletin of Mathematical Biology (1999) 61: 19-32)
71
72
Identifying predator-prey models (PhD Thesis)
C. Jost
Abstract
Ratio-dependent predator-prey models set up a challenging issue regarding their dynamics near the origin. This is due to the fact that such models are unde…ned at (0, 0).
We study the analytical behavior at (0, 0) for a common ratio-dependent model and
demonstrate that this equilibrium can be either a saddle point or an attractor for certain trajectories. This fact has important implications concerning the global behavior
of the model, for example regarding the existence of stable limit cycles. Then, we prove
formally, for a general class of ratio-dependent models, that (0, 0) has its own basin
of attraction in phase space, even when there exists a non-trivial stable or unstable
equilibrium. Therefore, these models have no pathological dynamics on the axes and at
the origin, contrary to what has been stated by some authors. Finally, we relate these
…ndings to some published empirical results.
Les modèles proie-prédateur du type ratio-dépendant posent un dé… concernant leurs
dynamiques proches de l’origine. Ceci est due au fait que ces modèles ne sont pas dé…nis
à (0, 0). Nous étudions le comportement analytique autour (0, 0) pour un modèle ratiodépendant simple et démontrons que cet équilibre peut être un point de sel ou un
attracteur pour certains trajectoires. Ce fait à des implications importantes concernant
le comportement globale du modèle, par exemple concernant l’éxistence de cycles limites
stables. Ensuite, nous prouvons formellement pour une classe générale de modèles du
type ratio-dépendant que (0, 0) tient son propre bassin d’attraction, même s’il y a un
équilibre non-trivial stable ou instable. Donc, ces modèles n’ont pas de comportements
dynamiques pathologiques sur les axes et à l’origine, contrairement aux énoncés de
certains auteurs. Finalement, nous comparons ces résultats avec quelques résultats
empiriques trouvés dans la littérature.
Jost, Arino, Arditi
5.1
Extinction in predator-prey models
73
Introduction
Continuous predator-prey models have been studied mathematically since publication
of the Lotka-Volterra equations. The principles of this model, conservation of mass and
decomposition of the rates of change into birth and death processes, have remained
valid until today and many theoretical ecologists adhere to these principles. Modi…cations were limited to replacing the Malthusian growth function, the predator per capita
consumption of prey or the predator mortality by more complex functions such as the
logistic growth, Holling type I, II and III functional responses or density-dependent
mortality rates.
The mentioned functional responses all depend on prey-abundance N only, but soon
it became clear that predator abundance P can in‡uence this function (Curds & Cockburn 1968; Hassell & Varley 1969; Salt 1974) by direct interference while searching or
by pseudo-interference (in the sense of Free et al. (1977)) and models were developed
incorporating this effect (Hassell & Varley 1969; DeAngelis et al. 1975; Beddington
1975). However, these models usually require more parameters and their analysis is
complex. Therefore, they are, on one side, rarely used in applied ecology and, on the
other side, have received little attention in the mathematical literature. A simple way of
incorporating predator dependence into the functional response was proposed by Arditi
& Ginzburg (1989) who considered this response as a function of the ratio N/P . Interesting properties of this approach have emerged that are in contrast with predictions
of models where the functional response only depends on prey abundance (e.g. Arditi
et al. (1991a), Ginzburg & Akçakaya (1992), Arditi & Michalski (1995)). Two principal
predictions for ratio-dependent predator-prey sytems are: (1) equilibrium abundances
are positively correlated along a gradient of enrichment (Arditi & Ginzburg 1989) and
(2) the ‘paradox of enrichment’ (Rosenzweig 1971) either completely disappears or enrichment is linked to stability in a more complex way. However, we will not discuss here
the general ecological signi…cance of this class of models but rather study a particular
mathematical feature of this model: the behavior around the point (0, 0) (where the
models are not directly de…ned) and its implications on global behavior. Interesting
dynamic behaviors such as deterministic extinction and multiple attractors can occur.
There are only few mathematical publications that study ratio-dependent models.
Many of them use logistic-type models where density dependence in the growth equation
is proportional to the ratio consumer/resource (e.g., the popular Holling-Tanner model
(Tanner 1975)). However, these models do not abide by the conservation of mass rule
(reproduction rate of predators is a function of the consumption rate, Ginzburg (1998)).
We are rather interested in ratio-dependent models that respect this conservation of
mass (or energy) as an important aspect of ecological modelling. This further reduces
the available literature on this class of models. Cosner (1996) developed ‡oor- and
ceiling functions to understand the behavior of complex systems that include temporal
variability, and ratio-dependent formulations proved to be more adapted to this kind of
study. Beretta & Kuang (1998) studied the in‡uence of delays on the stability behavior
of the non-trivial equilibrium.
Freedman & Mathsen (1993) studied conditions for persistence of a speci…c ratiodependent predator-prey model. They restricted their analysis to parameter values
Identifying predator-prey models (PhD Thesis)
74
C. Jost
that ensure that the equilibrium (0, 0) behaves like a saddle point. They based this
restriction on the assertion that attractivity of this trivial equilibrium is possible only
with parameter values for which the predator abundance P (t) increases without bound
as a function of time. In this paper, we will show that this assertion is erroneous
and we will reanalyze the general stability behavior of a typical ratio-dependent model
around the equilibrium (0, 0). Furthermore, we will give a formal proof (for a general
ratio-dependent model) that this point can become attractive for all initial conditions
sufficiently close to the predator axis, while the non-trivial equilibrium remains either
locally stable or becomes unstable. This gives rise to global behaviors that range from
global attractivity of the non-trivial equilibrium, coexistence of two different attractors
(each with its own basin of attraction) to global attractivity of the equilibrium (0, 0).
Extinction is a frequent outcome in simple laboratory predator-prey systems (Gause
1935; Luckinbill 1973) and biologists had to modify conditions in order to obtain (cyclic)
coexistence [e.g., spatial heterogeneities (Huffaker 1958) or viscous medium to slow down
the predators (Veilleux 1979)]. Since traditional predator-prey models predict cyclic
dynamics, extinction has been explained as the result of stochasticity occurring when
the trajectories come close to the axes. In this paper we show that, for some region in
the parameter space of a ratio-dependent model, multiple attractors can appear, one of
them being the origin. Therefore, extinction can be explained as a simple deterministic
process.
5.2
The model and its equilibria
A predator-prey system that incorporates conservation of mass and division of population rates of change into birth and death processes has the following canonical form:
dN
dt
dP
dt
= f (N )N
g(N, P )P
= eg (N, P )P
P
(5.1)
(5.2)
with prey abundance N (t) and predator abundance P (t), conversion efficiency e and
predator death rate . We will use the traditional logistic form for the growth function
f with maximal growth rate r and carrying capacity K :
f (N ) = r(1
N
).
K
The functional response g (prey eaten per predator per unit of time), that in general
depends on both prey and predator density, will be considered as a (bounded) function
of the ratio prey per predator,
g := g(
N
N/P
N
)=
=
1 + hN/P
P
P + hN
∀(N, P ) ∈ [0, +∞)2 \(0, 0)
(5.3)
with total attack-rate and handling time h. Note that the second equality is strictly
correct only for P > 0. In the case of P = 0 and N > 0 we can de…ne g(N, 0) := 1/h
(the limit of g(x) for x → ∞).
Jost, Arino, Arditi
75
Extinction in predator-prey models
ˆ = hN ,
In a …rst step we simplify this model by non-dimensionalisation. Let N
eK
h
rh
h
et
ˆ
,
,
,
and
.
In
these
new
variables
the
system
Pˆ = hP
R
Q
S
t
=
=
=
=
e2 K
e
e
e
h
becomes:
ˆ
SNˆ ˆ
dNˆ
N
(5.4)
P
= R(1
)Nˆ
S
dtˆ
Pˆ + SNˆ
dPˆ
SNˆ ˆ
(5.5)
P QPˆ
=
dtˆ
Pˆ + SNˆ
ˆ (0) = n0 , Pˆ (0) = p0 . For simplicity we will not write the hat ˆ
with initial conditions N
in the rest of this paper.
This system has at most three equilibria in the positive quadrant: (0, 0), (S, 0) and
a non-trivial equilibrium (n? , p? ) with
S(R + (Q 1)S )
R
S(1 Q) ?
n.
=
Q
n? =
p?
A simple calculation shows that n? is positive for all S < R/(1
Q < 1 and therefore ensures the positivity of p? .
Q), which implies
To see why (0, 0) is indeed an equilibrium (despite the fact that g is unde…ned in that
case) note that for any g that is a non-negative bounded function in its domain (such as
(5.3)) the right sides of system (5.1-5.2) become 0 at this point, which is the de…nition
of an equilibrium (boundedness of g is a sufficient condition, but not a necessary one).
Figure 5.1 shows the possible isoclines of the system. For S > R, the prey isocline is
a humped curve through the origin and the point (S, 0). For S < R, the denominator of
the prey isocline can become 0 for some N ∈ (0, S ). The part of the isocline that remains
in the positive quadrant becomes in this case a strictly monotonically descending curve
through the point (S, 0). The predator isocline is always a straight line through the
origin. See Arditi & Ginzburg (1989) for more details. While the cases of Figures
5.1a and 5.1c do not raise mathematical difficulties, the case of Figure 5.1b presents
interesting and unexpected mathematical properties that will be studied below.
5.3
Stability of the equilibria
The community matrix (Jacobian at the equilibrium) at the point (S, 0) is
R
1
0 1 Q
and therefore, if the non-trivial equilibrium exists (=⇒ Q < 1), this point is always a
saddle point.
The community matrix at (n? , p? ) has the form
R + S Q2 S
Q2
.
(Q 1)2 S
(Q 1)Q
76
Identifying predator-prey models (PhD Thesis)
C. Jost
Isop
Isop
Isop
Ison
Ison
Ison
Figure 5.1: The three general types of isoclines that can occur. (a): the non-trivial
equilibrium is stable and (0, 0) behaves like a saddle point (R = 0.5, Q = 0.3, S =
0.4). (b): both equilibria can be attractive or repelling, creating dynamics that
are illustrated in Figures 5.2 - 5.5. (c): the equilibrium (0, 0) is globally attractive
(R = 0.5, Q = 0.79, S = 3.0). The lines with arrows are examples of trajectories,
Ison is the prey isocline and Isop the predator isocline.
Applying the Routh-Hurwitz criterion shows that this equilibrium is stable whenever
Q Q2 + R
R
(5.6)
.
S < min
,
1 Q
1 Q2
2
Note that, if 1 RQ < Q 1 QQ+2 R (⇐⇒ R + Q < 1), then the non-trivial equilibrium is
always stable (if it exists). This is possible with two types of isoclines, Figures 5.1a
and 5.1b. The case of Figure 5.1b together with this condition (allowing arbitrarily low
stable equilibrium densities of both prey and predator) is particularly interesting in the
context of biological control where the interest is in non-trivial stable equilibria with
n? S. The non-trivial equilibrium in Figure 5.1a is also always stable (independently
of the above criterion), because its existence ensures that Q < 1, therefore, if S < R,
then S also ful…lls criterion (5.6). However, this case is less interesting because it requires
high predator densities to keep the prey density low.
At the equilibrium (0, 0) the community matrix cannot be calculated directly because
the ratio N/P is not de…ned at this point. To understand the stability behavior of this
point we must expand it on a whole axis by studying the transformed systems (N/P, P )
and (N, P/N ). Setting L := N/P , then we have the system
dL
= L(R + Q
dt
dP
SL
= (
dt
1 + SL
R
LP )
S
S
L(1 + L)
1 + SL
Q)P.
There are two equilibria on the L-axis, (0, 0) and ( S (SQ+QR R1) , 0). (0, 0) is a saddle point
for S < Q + R (eigenvalues of the community matrix are Q and Q + R S ), otherwise
it is attractive. The latter equilibrium has the eigenvalues
1 =
S(1
Q)
S 1
R
, 2 =
SQ + R
S + Q + SR
S 1
(R + Q)2
Jost, Arino, Arditi
77
Extinction in predator-prey models
and it is unstable whenever a non-trivial equilibrium exists.
Proof. Let S < 1. If the non-trivial equilibrium exists (S < R/(1
therefore the equilibrium is unstable.
Q)) then 1 > 0,
Now let S > 1. The existence of the non-trivial equilibrium ensures in this case that
R + Q > 1. Furthermore, S(SQ+QR R1) must be positive to be of interest, therefore
S >Q+R
(5.7)
and
2
=
1
S
>
=
1
1
(5.7)
S
1
1
S
1
(S (Q
((Q + R)(Q
(R(S
(R + Q)2 + SR
1) + (Q + R)
1+1
R
Q) + SR
(5.7)
(Q + R)) > 0.
This equilibrium is therefore unstable.
Finally, we need the stability behavior of (0, 0) for the system (N, M ) with M :=
P/N ,
dN
dt
dM
dt
NR
KS
= N R
S
K +S
M (S (N R (Q + R 1) S) + M (N R + S (S
=
S (M + S )
The community matrix at (0, 0) has the eigenvalues 1 = 1
point (0, 0) is therefore always unstable.
R
Q
R)))
.
Q, 2 = R and the
Summarizing we can conclude for the original system (N, P ) that for S < Q + R
the equilibrium (0, 0) behaves like a saddle point. For S > Q + R we have seen that
the system (N/P, P ) has an attractive equilibrium at its origin (0, 0). Interpreted in
the original state variables N and P this point can only be attained by a trajectory
for which ‘N goes faster to 0 than P ’. Below, we will discuss the existence of such
trajectories.
Freedman & Mathsen (1993), who studied in their paper the same model (5.4) and
(5.5), excluded the latter case (S > Q + R) from their persistence analysis of ratiodependent models by stating (p. 823) that “this implies that there are solutions (N (t),
P (t)) → (0, +∞) as t → ∞”. The following proposition proves that this statement
is erroneous.
Proposition 5.3.1. The system of equations (5.4-5.5) is ultimately bounded with some
bound independent of the initial values.
Identifying predator-prey models (PhD Thesis)
78
Proof. Let b, c > 0 such that
(R+b)2 S
4R
C. Jost
< c (for any b, such a c can be found).
(R + b)2 S
<c
4R
R
⇔ (R + b)2 4 c < 0
S
R 2
⇔ 0< N
N (R + b) + c
S
∀N
(5.8)
Therefore we have
d
(N + P )
dt
=
(5.8)
<
<
R 2
N
S
RN
QP
bN QP + c
d(N + P ) + c
with d := min(b, Q). So we can conclude that lim sup(N (t) + P (t)) dc . Note that we
t→∞
have N (t) + P (t) max(N (0) + P (0), dc ), ∀ t 0.
Freedman & Mathsen (1993) also point out that a general ratio-dependent model
can pose de…nition problems on the predator axis. However, if the functional response
is restricted to being positive and bounded (two properties not contested in ecology and
implicit in the model studied here), then (5.1) and (5.2) are perfectly well de…ned on
the whole positive quadrant [0, +∞)2 \(0, 0), and the analysis in this paper shows that
the behavior at (0, 0) has nothing abnormal that would justify its exclusion.
If the non-trivial equilibrium were unstable and the point (0, 0) a saddle point, then
we could construct easily a positive invariant set that contains these two equilibria and
apply the Poincaré-Bendixson theorem to prove the existence of a limit cycle. However,
the following proposition holds:
Proposition 5.3.2. For S < Q + R the non-trivial equilibrium (if it exists) is locally
stable.
This means that, if the non-trivial equilibrium is unstable, then S > Q+ R, implying,
as shown earlier, that (0, 0) is not a saddle point. This complicates considerably the
construction of the positive invariant set required to apply the Poincaré-Bendixson
theorem. We have not found such a set but do not exclude that it can exist.
Proof of proposition. a) For R + Q < 1 we have already seen above that all existing
non-trivial equilibria are stable.
b) For R + Q > 1, we have
R>1 Q
R = R(1 Q2 + Q2 ) > R(1
=⇒ R > (1 Q)(R + Q2 + RQ)
R
> (Q + R)(1 + Q) Q
=⇒
1 Q
R + Q(1 Q)
>Q+R>S
=⇒
(1 Q)(1 + Q)
Q2 ) + (1
Q)Q2
Jost, Arino, Arditi
Extinction in predator-prey models
79
and, according to criterion (5.6), the non-trivial equilibrium is stable.
Getz (1984) gave a proof of existence of a stable limit cycle for a ratio-dependent
model that only differed from the model used here by its prey growth function [ar/(bN +
r) c instead of r(1 N/K )]. He did not study rigorously the behavior at (0, 0), simply
stating that the isocline graph ‘demonstrates’ that it is a saddle point (as required by
the Poincaré-Bendixson theorem, since the origin is part of the positive invariant set
that he constructed). However, the general analysis in the next section applies also to
his system and it shows that (0, 0) can become attractive. His graphical interpretation
is therefore incorrect. It can be seen numerically that there are cases for which (0, 0)
becomes globally attractive instead of having a stable limit cycle around the non-trivial
equilibrium (as in Figure 5.5).
5.4
(0, 0) as an attractor
So far we have only shown that the equilibrium (0, 0) can be attractive for trajectories
where ‘N goes faster to 0 than P ’, but we do not know yet if this type of trajectory
really exists. In this section we will give a formal proof for this. This proof will be given
for any growth function f and any functional response g in the general form of system
(5.4-5.5),
dN
dt
dP
dt
= f (N )N
g(N/P )P
= g(N/P )P
QP,
with f and g having the following properties:
- f and g are continuous in R+ and both functions are bounded
- f (N ) < f (0)
∀N >0
- g(0) = 0, g0 (v ) exists and is positive for any v 0.
Proposition 5.4.1. Assume f (0) < g0 (0) Q (i.e. Q + R < S, in our system (5.45.5)). Then, any trajectory for which n0 is sufficiently small compared to p0 converges
to the point (0, 0).
Note that this proposition is a generalisation of a recent result by Kuang & Beretta
(1998).
Proof. Consider the system (N, L) with L :=
dN
= f (N )N
dt
dL
= f (N )L
dt
N
,
P
g(L)
N
L
(1 + L)g (L) + QL.
Identifying predator-prey models (PhD Thesis)
80
C. Jost
For L > 0 we have
d
L < f (0)L
dt
(1 + L)g (L) + QL
Because of our assumption f (0) < g0 (0)
the stronger inequality
Lg (L)>0
g(L)
L
L(f (0)
g(L)
+ Q).
L
Q we have for any ∈ (0, g 0 (0)
f (0) < g0 (0)
Since lim L→0
<
Q
Q
.
f (0))
(5.9)
= g0 (0) there exists some > 0 such that
|
g (L)
L
g0 (0)| < ∀ 0 < L < .
(5.10)
We can now conclude that
f (0)
(5.9)
g(L)
+ Q < g0 (0)
L
g(L)
L
=⇒
(5.10)
< 0
∀0<L<
(5.11)
d
L < 0 ∀ 0 < L < .
dt
Therefore, if there is some t0 with L(t0 ) , then L(t) ∀ t t0 and
d
L
dt
<0
=⇒ lim L(t) = 0.
t →∞
On this basis, we can further conclude that
d
N N (f (0)
dt
g(L) (5.11)
) < 0 ∀ t t0 .
L
Therefore lim t →∞ N (t) = 0. Finally, consider the equation
L(t) → 0 and g(0) = 0 there is some tω such that
d
P <0
dt
d
P
dt
= P (g (L)
Q), since
∀ t > tω
=⇒ lim P (t) = 0.
t→∞
This proves the proposition.
The condition Q + R < S is possible with isoclines as shown in Figures 5.1b and
5.1c. Examples of trajectories converging to the origin are shown in Figures 5.3–5.5. The
numerical simulations of the trajectories in these …gures were done using Mathematica
with the built-in high order adaptive step size procedure (the accuracy goal had to be
set higher than the default value to avoid numerical problems close to the origin).
Jost, Arino, Arditi
Extinction in predator-prey models
81
P(t)
Isop
Ison
N(t)
Figure 5.2: The non-trivial equilibrium is a global attractor and (0, 0) behaves like
a saddle point, S < Q + R. Parameter values are R = 0.5, Q = 0.79, S = 1.0.
5.5
Discussion
We saw that the equilibrium (0, 0) can behave in several ways depending on parameter
values. The following sequence of …gures illustrates these behaviors by steadily increasing parameter S while keeping parameters R and Q at …xed values. Figure 5.2 illustrates
the case for which it is a saddle point. All trajectories converge to the non-trivial stable equilibrium independently of the initial conditions (this equilibrium is therefore a
global attractor). Freedman & Mathsen (1993) derived for this case conditions that
ensure persistence of the predator-prey system. Figure 5.3 shows the case of having two
attractive equilibria, each with its own basin of attraction. The two basins were determined numerically by overlaying the phase space with a small scale grid, taking each
grid point as initial value and determining whether the simulation ends in (0, 0) or in the
non-trivial equilibrium. There must be a separatrix between these two basins. Figure
5.4 shows again a case with two basins of attraction, but the non-trivial equilibrium is
now unstable and we have a stable limit cycle. As was shown in the previous section we
cannot use the Poincaré-Bendixson theorem to prove the existence of this stable limit
cycle because the construction of a positive invariant set would require knowledge of the
analytic form of the separatrix. This …gure also shows that the limit cycles will be very
sensitive to stochastic in‡uences: random perturbations to the populations occurring
while the cycle is not far from the separatrix can bring the trajectory into the basin of
attraction of (0, 0), thereby causing extinction. Figure 5.5 shows the case when (0, 0)
becomes attractive for all positive initial conditions except the non-trivial equilibrium
itself. There is no formal proof of this global attractivity, and several trials with Dulac’s
criterion failed. Further increase of parameter S will make the non-trivial equilibrium
disappear and (0, 0) becomes (trivially) globally attractive (Figure 5.1c). The present
82
Identifying predator-prey models (PhD Thesis)
C. Jost
Figure 5.3: The non-trivial equilibrium is locally stable, but (0, 0) becomes also
attractive, S > Q + R. The light gray area is the basin of attraction of the nontrivial equilibrium, the dark gray area is the one of equilibrium (0, 0). Parameter
values are R = 0.5, Q = 0.79, S = 1.66.
mathematical analysis establishes that a general class of ratio-dependent models have
well de…ned dynamics on the axes and at the origin.
Figure 5.4: The non-trivial equilibrium is unstable and (0, 0) becomes attractive,
S > Q + R. There are two attractors, a stable limit cycle and (0, 0). Parameter
values are R = 0.5, Q = 0.79, S = 1.78.
Extinction of one or both populations in predator-prey systems have occupied ecologists since the classic experiments of Gause (1935), who tried to reproduce in the laboratory the cycles predicted by the Lotka-Volterra predator-prey equations. However,
Jost, Arino, Arditi
P(t) P(t)
Extinction in predator-prey models
83
Isop
Ison
N(t)
Figure 5.5: The equilibrium (0, 0) is a global attractor, S > Q + R. Parameter
values are R = 0.5, Q = 0.79, S = 1.85. There is no formal proof for the global
attractivity.
instead of the desired coexistence, the most frequent result was that the populations
(Paramecium sp. preyed upon by Didinium nasutum) went extinct either immediately
or after a couple of oscillations. Other researchers encountered the same problem (e.g.
Huffaker (1958), Luckinbill (1973)). By thickening the medium to reduce mobility of
the predator, Luckinbill (1973) obtained repeatedly several predator-prey oscillations
before extinction and Veilleux (1979) re…ned this technique to have …nally sustained cycles without extinction. He also did extensive experiments for various initial conditions
and detected two basins of attraction (his Figure 11) that are similar to those in our Figure 5.3. Since the classical predator-prey systems like Lotka-Volterra or more complex
ones with logistic growth and Holling type II functional responses cannot show deterministic extinction, these results have usually been explained by demographic stochasticity:
limit cycles bring the populations very close to 0 during the cycle and small stochasticities suffice to cause extinction. The model studied here can explain the extinction as
a deterministic result, with no need for stochasticity. The simultaneous existence of an
unstable non-trivial equilibrium and an attractive trivial equilibrium (0, 0) extends the
behaviors of this model from extinction after one simple oscillation, as brie‡y described
by Arditi & Berryman (1991), to extinction after a number of oscillations. Furthermore,
the technique of thickening the medium to stabilize the predator-prey interaction (Luckinbill 1973; Veilleux 1979) can be interpreted as reducing the attack rate (Harrison
1995) which, in the present ratio-dependent model, has a stabilizing effect. By varying
this parameter, the whole spectrum of observed behaviors (stable coexistence, sustained
oscillations, extinction after several cycles, immediate extinction) can be predicted, as
illustrated by Figures 5.2-5.5.
84
Identifying predator-prey models (PhD Thesis)
C. Jost
Acknowledgements
We thank Lev Ginzburg for emphasizing repeatedly the ecological interest of understanding extinction in predator-prey systems. This research was supported by the Swiss
National Science Foundation and by the French CNRS.
Chapter 6
Predator-prey theory: why
ecologists should talk more with
microbiologists
Christian Jost
Figure 6.1: David Ely Contois (1928–1988)
(Photo courtesy to the Department of Microbiology, University of Hawaii)
85
86
Identifying predator-prey models (PhD Thesis)
C. Jost
Abstract
Consumption of a resource by an organism is a key process in both microbiology and
population ecology. Recently, there has been a debate in population ecology about the
importance of organism density in functions describing this process. Actually, microbiologists have had this debate over the last 40 years. Reviewing their principal results
I show that even for the most simple systems there is no unique correct function to
describe consumption. Organism density in‡uences consumption to various degrees. I
conclude that, for predictions based on model simulation, one should use a pluralistic
approach, working with different models to identify robust predictions (that is, common to all studied models) and guide further research to understand model-speci…c
predictions.
La consommation de ressources par des organismes est un processus-clé à la fois en
microbiologie et en ecologie des populations. Un débat récent s’est ouvert en écologie
quant à l’importance de la densité des organismes dans les fonctions décrivant la consommation. En fait, cette question a également été débattue par les microbiologistes
durant les quarante dernières années. Par une revue de leurs principaux résultats, je
montre que, même pour les systèmes extrêmement simples, il n’existe pas une fonction universelle permettant de décrire correctement la consommation. La densité des
organismes in‡uence leur consommation à différents degrés. Je conclus que, lorsqu’on
fait des prédictions fondées sur des simulations, il est nécessaire d’avoir une approche
pluraliste en travaillant avec plusieurs modèles, de manière à identi…er les prédictions
robustes (c’est à dire celles qui sont communes à tous les systèmes étudiés), et à diriger
des études spéci…ques vers les points sur lesquels différents modèles font des prédictions
divergentes.
Jost
6.1
Ecologists talking with microbiologists
87
Introduction
Describing the consumption process in predator-prey interactions is a research topic
in population ecology since the early theoretical works of Lotka and Volterra. The
quantitative description of this process has faced several questions: does the instantaneous consumption depend only on food (prey) availability, or also on the consumers
(predators)? What function should be used in mathematical models? How should parameters for these functions be estimated? In this note I want to draw attention to
the work of microbiologists and its relevance to current debates in population ecology.
Microbiologists have often faced similar problems in describing the growth of bacteria
or protozoa on some substrate. Although hidden behind different names and notations,
several mathematical forms of consumption used in ecology have an equivalent microbiological growth function (Table 6.1). Most interestingly, there is also a twin (Contois
1959) to a model that has aroused a heated debate in ecology: the ratio-dependent
model introduced by Arditi & Ginzburg (1989). I will review in this note the results in
microbiology with respect to this twin model and discuss how they can help to …nd a
consensus in the ecological debate.
The functional response (prey eaten per predator per unit of time, Solomon 1949)
is traditionally considered to be a function of prey abundance only (prey-dependent,
Holling 1959b). However, predator density can also in‡uence individual consumption
rate, an effect that I will call predator dependence. Such predator dependence (usually
a decreasing functional response with increasing consumer density) has been observed
in many vertebrate and invertebrate species (reviews in Hassell 1978a and Sutherland
1996). A particularly simple way to include predator dependence has been proposed in
the ratio-dependent model where the functional response depends on the ratio prey density per predator density. This approach naturally predicts the experimentally observed
decreasing feeding rates with increasing predator densities and the positive correlations
between population abundances of producers and consumers observed along gradients
of productivity (see Arditi & Ginzburg 1989 and the review in Pimm 1991, p. 290).
Despite this empirical evidence supporting it, there is an on-going debate about the
validity of the ratio-dependent approach (Abrams 1994; Gleeson 1994; Akçakaya et al.
1995; Abrams 1997). The particular ratio-dependent growth function introduced in
microbiology much earlier by Contois (1959) has served there as an alternative to the
well known (prey-dependent) growth function of Monod (1942), a twin to the popular
Holling type II functional response (Table 6.1). Microbiologists have worked during the
last 40 years with Contois’ function, compared it to others or elaborated it further. I
have followed citations of Contois’ paper during this period of time and I will highlight
the results that are relevant for the on-going ecological debate. Brackets will indicate
in this review parallels and similar concepts in predator-prey theory.
Identifying predator-prey models (PhD Thesis)
88
6.2
C. Jost
A short historical perspective
The introduction of Monod’s growth function in 1942,
g(s) = max
s
,
Ks + s
(6.1)
(parameters are explained in Table 6.1) together with its mathematical handyness and
strong experimental and theoretical/methodological support, was a major breakthrough
in the mathematical description of bacterial growth. Microbiologists have applied Monod’s model to the description of monospeci…c organisms growing on a homogeneous
substrate in batch and chemostat cultures (reviews in Jannasch & Egli (1993) and
Fredrickson (1977)). These chemostats can be described by:
ds
= D(s0
dt
dx
= g(s)x
dt
s)
D(x
Y g(s)x
x0 )
(6.2)
with yield Y , dilution rate D, in‡owing substrate concentration s0 , and in‡owing organism concentration x0 (usually equal to 0). The last three parameters can be controlled
entirely by the researcher. [These equations correspond to predator-prey equations
with constant prey immigration. However, there is a subtle difference between microbiological and ecological modeling: while microbiologists start with the growth function
(ecologists call it numerical response) and consider the substrate uptake function to
be proportional to it (yield Y ), ecologists often start with the functional response and
consider the numerical response to be proportional to the functional response (with
conversion efficiency e = 1/Y , see equations 6.2 and 6.4 below). This difference is
nevertheless of little importance for the qualitative results that will be discussed below.]
Table 6.1: References to the same model in ecology and microbiology. s is prey
density or substrate concentration, x is predator density or density of organism
that grows on s, a is predator attack rate, h is handling time, max is the maximum
growth rate, Ks the Michaelis-Menten or half saturation constant, total searching
efficiency and c, m are empirical positive constants.
functional response
reference in
reference in
or growth rate
ecology
microbiology
as
s sb
Holling (1959) I
Blackman (1905)
asm s sb
with upper limit
as
= max Kss+s
Holling (1959) II
Monod (1942)
1+ahs
Ivlev (1961)
Teissier (1936)
a(1 e cs )
asm
Real (1977)
Moser (1958)
1+ahsm
s/x
s
= x+hs
Arditi & Ginzburg
Contois (1959)
1+hs/x
(1989)
s 1
Hassell & Varley
Ashby (1976)
max Ks +s x
(1972) (special case)
Jost
Ecologists talking with microbiologists
89
Despite its initial success, there were experimental results that could not be explained with Monod’s function. At …rst, these were attributed to apparatus effects such
as incomplete mixing or growth on chemostat walls (e.g., Herbert et al. 1956). Contois
(1959) was the …rst to suggest and to present experimental results that the half saturation ‘constant’ Ks is in fact not a constant (estimates of this ‘constant’ varied up
to three orders of magnitude, see Jannasch & Egli 1993) but that it is proportional to
in‡owing substrate concentration, Ks = ks0 . Together with the occurrence of mass balance ( ds
+ Y dx
= 0 in system (6.2), which suggests the relation (x x0 ) = x = Y (s0 s))
dt
dt
model (6.1) changes to
g(s, x) =
max s
max s
max s/x
=
=
k(x/Y + s) + s
(k/Y )x + (k + 1)s
(k/Y ) + (k + 1)s/x
(6.3)
which is a particular case of a growth function that depends on the ratio substrate per
organism s/x.
Curds & Cockburn (1968) gathered experimental evidence that the growth rate of
protozoa feeding on bacteria is a decreasing function of the protozoan concentration.
[The same phenomenon in ecology is what I termed predator dependence]. This is in
contrast to Monod’s function which predicts that the growth rate should be independent
of organism concentration. This negative dependence of the growth rate on organism
concentration was also con…rmed by Aiba et al. (1968), Fayyaz et al. (1971) and
Wilhelm (1993), and was usually explained as the result of accumulation of metabolic
byproducts that inhibit growth.
Monod’s function (6.2) also predicts that effluent substrate concentration in chemostats should only depend on dilution rate D and be independent of in‡uent substrate
concentration s0 . [This is equivalent to the vertical predator isocline in Lotka-Volterra or
Rosenzweig-MacArthur predator-prey systems; Figure 6.2]. This prediction was tested
by varying dilution rates and in‡uent substrate concentration, letting the chemostat
reach steady state and measuring then effluent substrate concentration s. Monod’s prediction was con…rmed for pure cultures growing on glucose (Grady Jr. et al. 1972), but
the results consistently diverged from this prediction when working with mixed cultures
(e.g., in wastewater treatment or fermentation processes) (Grady Jr. et al. 1972; Grady
Jr. & Williams 1975; Elmaleh & Ben Aim 1976; Daigger & Grady Jr. 1977). In the
latter, the out‡owing substrate concentration was proportional to in‡owing concentration, as predicted from the chemostat equations (6.2) with Contois’ function (6.3). [It
corresponds in ecology to the prediction that the prey equilibrium in a ratio-dependent
predator-prey system is proportional to prey carrying capacity].
The …rst approach to reconcile theory and experiment was to introduce intermediate
models that contain both Monod’s and Contois’ functions as special cases, e.g., Roques
et al. (1982) and Borja et al. (1995),
g(s, x) =
s
.
Ks + s + cx
[This form was introduced independently in ecology by DeAngelis et al. (1975) and
by Beddington (1975).] Kargi & Shuler (1979) proposed another intermediate function
that attempted to unify Monod’s, Contois’, Teissier’s and Moser’s growth functions
90
Identifying predator-prey models (PhD Thesis)
C. Jost
in the context of chemostats. However, experimentalists rarely use these intermediate
functions because of the effort required to estimate the additional parameter, while
theoreticians do not like them because of the considerably more complicated analytical
expressions. [The DeAngelis-Beddington function encounters the same fate in ecology.]
The second approach was to confront data directly with different functional forms of
the growth rate, by …tting either the dynamic model (6.2) to time series data of substrate
and organism abundances, or by …tting direct measurements of the growth rate as a
function of organism and substrate. Model selection was then based on the goodness-of…t criterion. Table 6.2 lists these studies together with the tested models (the best …tting
model in capitals). Often Contois’ function …tted best, but the differences in goodnessof-…t were usually small. In an interesting application of catastrophe theory to the
analysis of a protozoan system Bazin & Saunders (1978) found that, if the ratio prey
per predator is taken as the critical variable, then “a comparatively simple mechanism
can account for the observed behaviour”.
In summary, the experimental results illustrate that Contois’ model was successfully
used in the context of mixed cultures on multicomponent substrates such as wastewater
treatment, fermentation processes or biogas production from manure. While most of its
support is empirical, Contois’ equation has also been derived by mechanistic reasoning
(Fujimoto 1963; Characklis 1978) based on enzyme kinetics or saturation kinetics applied
to mass transfer limited growth. The empirical evidence suggests a mixed result: Contois’ model has only been contradicted by experimental results for monospeci…c cultures
growing on pure medium. In all other cases (mixed medium or several species/strains
present) it was rather Monod’s model that should have been rejected. [Similarly in ecology, where the particular predictions of prey-dependent food chains could only be found
in protozoan laboratory systems with pure strains for prey and predator species (Kaunzinger & Morin 1998)]. Monod’s function remained nevertheless the predominant one in
the microbiological literature (e.g., Barford & Hall 1978; Jannasch & Egli 1993). Why
so? I think this is due mainly to historical reasons (Monod published before Contois and
he is by far a more in‡uential biologist), but the current teaching of microbial modeling
also bears its share for generally presenting Monod’s model as the basic model, without
mentioning alternatives. [In the same way, introductory ecology books only mention the
prey-dependent Holling type I, II and III models, rarely do they present any alternative
functions]. Modeling efforts often start with Monod’s model and stop upon obtaining a
reasonable …t, without testing whether an alternative model can explain the results as
well or better (the papers cited in Table 6.2 show that most consumption functions …t
qualitatively correctly to the experimental data). Like a vicious circle, Monod’s model
con…rms itself without giving another model the chance to be tested as well.
6.3
Lessons for ecology
Microbiological experiments are usually done in well controlled laboratory situations,
following the observed processes with precise measurement techniques. In contrast, …eld
ecology has to cope with various stochastic in‡uences and unprecise census techniques.
Therefore, empirical validation of a model from data is higher valued in microbiology
Jost
Ecologists talking with microbiologists
91
than in ecology where model rejection by reason of data incompatibility is rare. However, Table 6.2 also lists results where several growth functions …tted equally well to
the same (microbiological) data. Therefore, as a …rst lesson, if model selection based on
goodness-of-…t is ambiguous even with microbiological data, then model selection based
on ecological …eld data cannot be expected to be any better. I would speculate that
this is one of the reasons why many ecologists consider mechanistic underpinning an
essential part for the validity of a model, even though the models, by their very simplicity, cannot be much more than a phenomenological account of the observed biological
processes. Furthermore, “one scientists’ mechanism is another scientists’ phenomenon”
(Pimm 1991). This quote best summarizes that the very de…nition of mechanistic is
controversial and ‘mechanisms’ are rather a methodological support than a true representation of what is happening in a population. Ratio dependence was also introduced
as an empirical way to include predator dependence, only recently have mechanistic
derivations been developed (Poggiale et al. 1998; Cosner et al. in press).
Let us now interpret the ecological meaning of the cited results in microbiology.
The most often tested prediction of Monod’s approach has been that the effluent substrate concentration in chemostats should be independent of in‡uent substrate concentration. This prediction is equivalent to the one that in predator-prey systems with
prey-dependent functional response G(s),
ds
s
= r(1
)s G(s)x
dt
K
dx
= eG(s)x x,
dt
(6.4)
the equilibrium prey-density is independent of carrying capacity K (that takes the role
of in‡owing substrate concentration) and that the predator isocline is vertical (as in
Lotka-Volterra or Rosenzweig-MacArthur predator-prey systems). See Figure 6.2a and
b for examples in chemostats or predator-prey systems. This prediction from Monod’s
function has been con…rmed only in the case of growth of a single organism type on
a single substrate within the constant environment of chemostat or batch cultures.
Every deviation from these conditions leads to effluent substrate concentration being
proportional to in‡owing concentration. This result can only be explained with a slanted
predator isocline (see Figure 6.2c to f), and it is also an essential prediction of the
ratio-dependent model. The second series of results in batch and continuous cultures
shows that the growth function is a decreasing function of predator density whenever
heterogeneities occur either in consumer species composition or in the substrate, again
suggesting that the predator isocline should be slanted. Such a slanted predator isocline
seems to be the rule rather than the exception and ratio-dependence with its isocline
through the origin (Figure 6.2e and f) can be considered a parsimonious way to model
it.
The consideration of a vertical predator isocline lies also at the base of two so-called
paradoxes: the paradox of enrichment (Rosenzweig 1971) that predicts that richer systems (high K ) should be less stable, and the paradox of biological control (Arditi &
Berryman 1991) that predicts that biologically controlled pests should have unstable
dynamics. These paradoxes are often resolved by creating more complex models with
92
Identifying predator-prey models (PhD Thesis)
C. Jost
Table 6.2: Collection of studies that compared Contois’ function quantitatively
with other functions (1-4, the tested models are noted in the third column with
the best …tting model, if there was one, in capitals) or studies that did a simple
model validation with Contois’ function (5-12). References are: (1) Chiu et al.
(1972), (2) Morrison et al. (1987), (3) Dercová et al. (1989), (4) Wilhelm (1993),
(5) Fujimoto (1963), (6) Goma & Ribot (1978), (7) Kristiansen & Sinclair (1979),
(8) Pareilleux & Chaubet (1980), (9) Lequerica et al. (1984), (10) Tijero et al.
(1989), (11) Bala & Satter (1990), (12) Ghaly & Echiegu (1993), (13) Benitez et al.
(1997).
ref. studied system
tested functions
(1)
microbial sewage
Moser, Monod,
Contois
(2)
nutrient limited phytoplankton growth
Monod, Contois, logistic
(3)
growth and glucose consumption of yeast
Contois,
Monod
(4)
protozoan feeding rate on bacteria
Contois, Monod
and 9 others ?
(5)
bacterial growth on yeast
(6)
hydrocarbon fermentation (with Contois’
function at low substrate concentrations, a
modi…ed one at higher concentrations)
(7)
production of citric acid in single stage continuous culture
(8)
aerobic cultures of apple fruit cells
(9)
anaerobic fermentation of rice straw
(10) anaerobic digestion of glucose and sucrose
(11) substrate degradation and biogas production from cattle waste
(12) continuous ‡ow no-mix anerobic reactor of
daily manure
(13) aerobic degradation of olive mill wastewaters
? best
…ts were obtained with functions that are sigmoid with respect to
substrate concentration, either Contois or Monod type
Jost
Ecologists talking with microbiologists
Predator-Prey
Chemostat
(a)
(b)
(c)
(d)
predator
organism
93
(f)
(e)
substrate
prey
Figure 6.2: Typical isoclines in chemostats (left) and predator-prey sytems (right)
with prey-dependent Holling II growth functions (a,b), DeAngelis-Beddington type
growth functions (c,d) and ratio-dependent growth functions (e,f). Prey isoclines
are long-short dashed, predator isoclines are short dashed. Straight lines represent
typical trajectories.
additional state variables or additional parameters (e.g., McCauley et al. (1988), Scheffer & de Boer (1995)). These additions rapidly lead to analytically intractable models
and discourage any further investigation. Instead of …xing the theory in an ad hoc manner to accommodate to particular cases, we might alternatively ask whether the ‘basic
model’ itself might have ‡aws and start in a modeling framework that inherently has a
slanted predator isocline. After all, “modeling philosophies . . . should be treated in the
same way as models – retained only as long as they assist progress” (Nisbet & Gurney
1982).
To continue this line of thought, consider the ecological equivalent to the cited …nding
that the growth function decreases with increasing consumer density (i.e., the question
of predator dependence). There is ample empirical evidence that interference between
foraging predators (a likely cause for predator dependence) is a frequent phenomenon in
invertebrates (review in Hassell 1978a) and in vertebrates (review in Sutherland 1996).
However, this reported interference usually just triggers the remark that a functional
response of the Hassell-Varley type (Hassell 1978a) would be most appropriate to model
the system. However, due to a lack of experimental data to parameterize this model
and because of analytical reasons, one is compelled to return to simple prey-dependent
types that do not take this predator dependence into account. Ratio dependence offers
a simple theoretical framework that inherently contains predator dependence either for
94
Identifying predator-prey models (PhD Thesis)
C. Jost
food chains (Arditi & Ginzburg 1989) or for whole food webs (Arditi & Michalski 1995),
while keeping models as simple and as tractable as those coming from the modeling
frameworks based on Holling type functional responses.
Some authors continue arguing that the ‘recent focus on ratio dependence is unfortunate’ (Abrams 1997; Murdoch et al. 1998), de‡ecting attention from more general
forms of predator dependence. I think that, on the contrary, ratio-dependent theory
has enhanced the status of predator dependence. It provides a simple mathematical
framework to test whether strong predator dependence changes predictions that were
originally derived from prey-dependent concepts, and to guide further research when
these predictions are not ‘robust’ against predator dependence such as the two paradoxes mentioned above. While modeling frameworks based on prey-dependent interactions can be linked to top-down mechanisms (Arditi & Ginzburg 1989) and account for
cycling systems (Rosenzweig 1971), the ratio-dependent approach includes elements of
top-down and bottom-up regulation (Arditi & Ginzburg 1989) and offers the possibilitiy
of deterministic extinction (Jost et al. 1999).
Ecology can pro…t from all these modeling frameworks. The reviewed results show
that natural systems contain in general predator dependence, but they do not tell which
of the modeling frameworks is a better approximation. Using both and comparing their
predictions can serve to guide further research in case of different predictions, while
similar predictions give con…dence having found ‘robust’ features of the studied systems.
Acknowledgements
I thank R. Arditi for initially pointing out the equivalence between Contois’ model
and ratio dependence and for supporting this study. I thank P. Inchausti for helpful
discussions and a careful reading of the manuscript. This research was supported by
the Swiss National Science Foundation and by the French ‘Programme Environnement,
Vie et Société’ (CNRS).
Chapter 7
Identifying predator-prey processes
from time-series
Christian Jost, Roger Arditi
95
96
Identifying predator-prey models (PhD Thesis)
C. Jost
Abstract
The functional response is a key element in predator-prey models as well as in food
chains and food webs. Classical models consider it as a function of prey abundance only.
However, many mechanisms can lead to predator dependence, and there is increasing
evidence for the importance of this dependence. Identi…cation of the mathematical form
of the functional response from real data is therefore a challenging task. In this paper
we apply model-…tting to test if typical ecological predator-prey time-series data, that
contain both observation error and process error, can give some information about the
form of the functional response. Working with arti…cal data (for which the functional
response is known) we will show that with moderate noise levels, identi…cation of the
model that generated the data is possible. However, the noise levels prevailing in real
ecological time-series can give rise to wrong identi…cations. We will also discuss the
quality of parameter estimation by …tting differential equations to this kind of timeseries.
La réponse fonctionnelle est un élément clé dans les modèles proie-prédateur, ainsi
que dans les modèles de chaînes et de réseaux trophiques. Dans les modèles les plus classiques, la réponse fonctionnelle dépend uniquement de l’abondance des proies. Toutefois, divers mécanismes peuvent également faire intervenir l’abondance des prédateurs.
Des données empiriques de plus en plus nombreuses suggèrent que celle-ci joue un rŹle
primordial. Il est donc important d’identi…er la forme mathématique de la réponse
fonctionnelle. Dans le présent article, nous utilisons les techniques d’ajustement de
modèle pour déterminer si cette identi…cation est possible sur des données écologiques
réelles, comportant un "bruit" dû aux erreurs de mesure et aux stochaticités environnementales et démographiques. Sur des données arti…cielles, créées avec une réponse
fonctionnelle connue, nous montrons qu’avec un "bruit" modéré l’identi…cation du modèle ayant généré les données est possible. Toutefois, les niveaux de "bruit" typiques que
l’on rencontre dans les séries temporelles en écologie peuvent mener à des identi…cations erronnées. Nous discutons également de la qualité des estimations de paramètres
obtenus par ajustement d’équations différentielles à de telles séries teporelles.
Jost, Arditi
7.1
Identifying predator-prey models I
97
Introduction
Finding the functional relationship between observed data is one of the major tasks in
ecology. Often several functional forms, that are based on different assumptions about
the dominant mechanisms at work, are available. Fitting these functions to the data
and applying goodness-of-…t as a criterion to select the best model is then used to detect
the dominant mechanism for the particular system from which the data were obtained.
One particular application of this concept is to test dynamic predator-prey models against predator-prey time-series data. Harrison (1995), for example, reanalysed
Luckinbill’s (1973) classical protozoan data and …tted them to 11 different (continuous)
predator-prey models. He assumed that the data contain only noise due to observation
error (measurement error), which leads to …tting the whole trajectory of the predatorprey system to the time-series (termed observation error …t by Pascual & Kareiva 1996).
Unfortunately, his statistical analysis did not take into account the number of parameters. It is therefore not too surprising that a rather complicated model with 11 parameters …tted the data best. Carpenter et al. (1994) …tted (discrete) predator-prey models
to phyto- and zooplankton time-series from North American freshwater lakes to test
whether the predation process depends signi…cantly on predator density. Their analysis
was designed to treat data that contain noise due to observation error and noise due to
process error (environmental or demographic stochasticity), …tting such that the prediction one time step ahead is minimized (termed process error …t by Pascual & Kareiva
1996). To avoid any assumptions about the presence or absence of higher predation
on the predators they …tted only the prey equation, using the predator data as input.
While they took the number of parameters into account, they did not justify the use of
discrete models (with the time step being the time between measurements) to describe
a system showing the characteristics of a continuous system.
However, does a better …t of one model compared to another one always imply that
its functional form represents the actual processes at work more accurately? There
exist, for example, simple algebraic differential equations that can …t perfectly to any
…nite time-series (Rubel 1981). It can also happen that very different models …t equally
well to the same data (Feller 1939). A slightly better …t of one of these models could
be an artifact of the time-series being one particular realisation of an ecological process
with all its random in‡uences. Another realisation (replicate) might give a very different
result. Therefore, the reliability of goodness-of-…t to determine the functional form of
a process from time-series data should be tested in advance, e.g., with arti…cal data for
which the functional form is known.
Carpenter et al. (1994) were aware of this problem and they tested their method with
arti…cial data that they created with parameters characteristic for their limnological
system. In this article we will perform a similar analysis for a larger range of ecological
systems. We will test whether predator-prey time-series that represent the (continuous)
dynamics of a stable focus contain sufficient information to detect if predator density
in‡uences the predation process strongly enough to in‡uence the dynamics of the system.
Such time-series typically contain noise due to observation error and noise due to process
error. The …tting techniques will include observation error …t (Harrison 1995) and a
modi…ed process error …t (Carpenter et al. 1994) that predicts s-steps ahead instead of
98
Identifying predator-prey models (PhD Thesis)
C. Jost
simply one-step ahead. The idea is to predict over a time range where non-linear effects
become detectable. The determination of s will be based on the arguments developped
and justi…ed in Ellner & Turchin (1995).
We will work with very simple predator-prey models whose purpose is not to describe
the data perfectly well but rather to describe them in a qualitatively correct manner.
Simplicity in the description of the key processes (growth, death) is essential in models of
more complex food chains or whole food webs, where the number of parameters becomes
a limiting factor for analysis and parametrization. Although such complex models are
not the subject of this article, it is with this purpose in mind that we deliberately
consider simple predator-prey models.
Based on the principles of mass conservation and decomposition of the dynamics of
a population into birth and death processes, the canonical form of such a predator-prey
model is
dN
= f (N )N g(N, P )P =: FN (N, P )
dt
dP
= eg (N, P )P P =: FP (N, P )
(7.1)
dt
where N and P are the abundances of prey and predator respectively, e the conversion
efficiency and the death rate of the predator in the absence of prey. The key processes
are the prey growth function f and the link between prey and predator, the functional
response g (prey eaten per predator per unit of time, Solomon 1949). The latter represents the predation process. We will test if model …tting can reveal whether g
is approximately a function of prey abundance only (g = g(N ), as it is the case in
traditional functional response models, e.g., Lotka 1924 or Holling 1959b) or if g also
depends signi…cantly on predator abundance. Such predator dependence in‡uences the
stability of predator-prey systems (DeAngelis et al. 1975; Murdoch & Oaten 1975) and
the response of the prey equlibrium to an enrichment of the system (Arditi & Ginzburg
1989). Its detection from natural predator-prey time series is therefore a challenging
task.
Introducing explicit predator dependence, g = g(N, P ), as was done by DeAngelis
et al. (1975), normally increases the complexity of the function g, making it difficult
to compare with the simpler prey-dependent form g = g(N ). A special case of a simple
predator-dependent function was suggested by Arditi & Ginzburg (1989), assuming that
g = g(N/P ). Models of this type are equally simple as prey-dependent models and can
therefore be directly compared with them. However, ratio dependence represents only
one particular case of predator dependence, and the only reason to favour it against other
predator-dependent functions is its simplicity. This ratio-dependent functional response,
like other predator-dependent functional responses, but in contrast to prey-dependent
functional responses, leads to the observed correlated equilibria of prey and predators
along a gradient of richness (Arditi & Ginzburg 1989; Mazumder 1994; McCarthy et al.
1995). The issue of ratio-dependence is currently subject to some debate (Abrams 1994;
Sarnelle 1994; Akçakaya et al. 1995; Abrams 1997; Bohannan & Lenski 1997; Hansson
et al. 1998).
In this article, we do not address the question of the ecological signi…cance of one
or the other model. We will merely attempt to answer the question of whether typical
Jost, Arditi
Identifying predator-prey models I
99
predator-prey time-series can help evaluating the importance of predator dependence.
In particular, we want to analyse the dynamics of predator-prey systems with low initial
conditions and whose trajectory reaches a stable, non-trivial equilibrium (both populations coexisting) after one or two large amplitude oscillations. Such dynamics are
considered characteristic for seasonal dynamics of phyto- and zooplankton in freshwater
lakes of the temperate zone (Sommer et al. 1986) or for chemostat and batch culture
experiments with protozoa (another source of published time-series data, e.g., Gause
et al. 1936, Luckinbill 1973). Typically, such time-series are short (about 20 data
points per season in lakes, 10-50 data points with protozoa), and may have considerable
observation and process error. Differential equations seem the adequate tool to describe
these systems since there are overlapping generations and large numbers of individuals.
Using a simulation approach, we will generate arti…cial time-series (termed pseudo-data
by May 1989) with a prey-dependent and a ratio-dependent functional response of the
same simplicity to which we will add process and observation error. Regression techniques will then be applied and we will test whether the best-…tting model is indeed
the one that created the data. A by-product of this kind of identi…cation is the computation of the actual model parameters. We will analyze the quality of these estimates
(value, standard deviation) to test the power of the regression method for parameter
estimation.
7.2
The alternative models
Building on the canonical form (7.1) we use a standard logistic growth for the reproduction function f ,
f (N ) = r(1
N
),
K
with maximum growth rate r and carrying capacity K . Two models are chosen for
the functional response, a prey-dependent one and a predator-dependent one, that have
both the same number of parameters. We chose the classical Holling type II model on
the one side and a ratio-dependent model (Arditi & Michalski 1995) on the other side,
aN
1 + ahN
g(N, P ) →
N/P
,
1 + hN/P
where a is the searching efficiency, h the handling time and some kind of ‘total’
predator searching efficiency. We selected the Holling type II form rather than equally
plausible alternatives such as the Ivlev functional response (Ivlev 1961) simply because it
is more widely used in ecology as well as in microbiology (Monod 1942). The particular
form of the predator-dependent functional response closely resembles the Holling type
II function, thus making direct comparison between the two models possible. This form
also tends to be regarded as the standard form of a ratio-dependent functional response
in mathematical studies (Freedman & Mathsen 1993; Cosner 1996; Kuang & Beretta
1998) and it is known in the microbiological literature as Contois’ model (Contois 1959).
Despite their structural difference, the two models can produce very similar temporal
dynamics. This is illustrated in Figure 7.1: time-series were created with both models
100
Identifying predator-prey models (PhD Thesis)
C. Jost
(with parameters corresponding to a stable focus), adding stochastic noise and observation error, and then …tting both models crosswise to these time-series (see the next
section for the details of these methods). It can be seen that both models …t very well
to the data created by the other model. A good …t alone is therefore a poor indicator
whether the used model correctly describes the processes that generated the data.
(b)
abundances
abundances
(a)
time
time
Figure 7.1: Examples illustrating that each model (the prey-dependent and the
ratio-dependent one) can approximate satisfactorily time series data that were
created by the other model. (a) Ratio-dependent model …tted to prey-dependent
data. (b) Prey-dependent model …tted to ratio-dependent data. Diamonds represent the prey time-series and stars the predator time-series. See the text for
further details.
7.3
7.3.1
Materials and methods
Arti…cial time-series
For this analysis to be valid for many different predator-prey systems, the pseudo-data
must be generated with widely differing parameter values. Possible parameter values
must abide to ecological and dynamical constraints. Such a constraint applies to the
conversion efficiency e that should be within the interval (0, 1) if abundances of both
prey and predator are measured in biomass (the usual case in freshwater studies). Parameters K and h can be chosen arbitrarily since they depend entirely on the time and
weight scales that are used. Given these three parameters, we can …nd intervals for
the remaining parameters by the requirement de…ned above: existence of a non-trivial
stable equilibrium reached with oscillations. Within these intervals, the parameters are
chosen randomly. Initial values of prey and predator abundances are then chosen two
to ten times below their equilibrium abundances. Such a randomly created parameter
set (with initial values) is retained only if the following properties are respected: (1)
prey and predator equilibria do not differ by more than a factor of 100, and (2) the
deterministic trajectories of prey and predator show at least two distinct oscillations
before reaching the equilibrium. The simulation time T is set in order to have these two
oscillations. These …nal criteria assure an at least twofold variation in predator abundance (that is essential for model identi…cation) and keep prey and predator abundances
on comparable scales (see Figure 7.1 for two examples).
Jost, Arditi
Identifying predator-prey models I
101
For each functional response, 20 such parameter sets were created. In analogy with
the replicates of a typical ecological experiment, we created with each parameter set 5
replicate time-series by numerical integration of the stochastic version of the differential
equations (7.1),
Nt+t = Nt + FN (Nt , Pt )t + p Nt N,t t
Pt+t = Pt + FP (Nt , Pt )t + p Pt P,t t,
(7.2)
with N,t and P,t being random normal variates with mean zero and variance one,
t := T/500 and p the process error level. This stochastic process was sampled at
20 equal time steps and a lognormally distributed observation error (with coefficient
of variation CV ) was incorporated by multiplication with the exponential of a normal
variate with mean zero and variance log(1 + CV 2 ). With this formulation, both process
error and observation error are of a multiplicative type as suggested to be typical for
natural populations (Hilborn & Mangel 1997; Carpenter et al. 1994). Time-series with
two noise levels were created, with CV and p both set to 0.05 and to 0.1. The …rst
case is comparable to protozoan laboratory data and the latter to data from freshwater
plankton experiments (Carpenter et al. 1994). This makes a total of 400 data sets (2
models * 20 parameter sets * 5 replicates * 2 noise levels).
7.3.2
Error functions
The key part in …tting a model to data is the formulation of the function to be minimized.
Depending on the stochastic elements in the data (process and/or observation error),
the error function must be chosen accordingly. Ecological data have usually both types
of error. However, statistical methods that take both into account are rare and little
is known in the case of nonlinear regression. The usual practical solution is therefore
to neglect one of the errors and to develop the error function for the other (Pascual
& Kareiva 1996). We will follow this approach, but also test two error functions that
claim to be able to take both errors into account.
For ease of notation, consider a simple autonomous differential equation y˙ = f (y )
with time-series data (ti , Yi )1im , where ti is the time at which the population y is
observed to have density Yi , and m is the number of data points. [For predator-prey
models y has to be replaced by the pair (N, P ) and adaptations for this case that
are not obvious in the development below will be noted in brackets]. Let y(ti ) be the
deterministic solution of the differential equation at time ti and yˆi the (unknown) real
population density at time ti . If the data have only observation error, then there is only
one initial condition, y(t0 ) = yˆ0 , that is treated as a free parameter. If there is only
process error, then the initial conditions are different for each consecutive data point
and are de…ned as the data value s steps previously, y(ti s ) = Yi s (see Figure 7.2). s is
chosen as the smallest value for which the autocorrelation in the time-series is below 0.5.
Ellner & Turchin (1995) had developped this method to choose s on empirical grounds
and argued that nonlinear patterns can be detected more reliably with this s-step ahead
prediction than by the traditional one-step ahead prediction. In our arti…cial data, s
102
Identifying predator-prey models (PhD Thesis)
C. Jost
always took the value 2. With these notations, the process can be written as
observation error only
yˆi = y(ti , )
Yi = yˆi ω
y(t0 ) = yˆ0
process error only
yˆi = y(ti , ) (ti , f, )
Yi = yˆi
y(ti s ) = Yi s
where is the vector of model parameters. We will suppose that the observation error
ω in the densities is of a multiplicative type (lognormal), as used in Carpenter et al.
(1994) and Hilborn & Mangel (1997), with a constant coefficient of variation CVω . The
(accumulated) process error i depends in general on the time interval ti = ti ti 1
and on the dynamics f over this interval. However, for many ecological time-series the
interval ti is constant and we further simplify by ignoring the effect of the dynamics
of f . Thus, process error i will be considered to be a lognormal variate with
constant CV (Carpenter et al. 1994). These two lognormal errors are considered to be
exponentials of two normal variates with expectation 0 and variance k2 = log(CVk2 + 1),
k = ω, .
Observation error fit
y t4)
Y4
(a)
t
t4
Process error fit
y t4)
Y4
(b)
t
t4
Figure 7.2: The …tting procedure changes with different types of error in the data.
(a) Observation error only. The whole trajectory is …tted to the data, treating
initial conditions as parameters. (b) Process error only. The best approximation
is to …t from one point to the next (one-step ahead prediction).
The log-transformed data Yi are therefore Gaussian with expectation log(yˆi ) and
standard deviation k . We de…ne the residuals
di,k = log(Yi )
log(yk (ti )),
k = ω, .
(7.3)
Jost, Arditi
Identifying predator-prey models I
103
The index k refers to the values yk (ti ) which are computed with different initial conditions according to the type of error. [Note that the residual di,k is a vector for vectorvalued yk (ti )]. Assuming now only one type of error, the error function to be minimized
becomes:
Xk2
=
m
2
X
di,k
i=1
k2
k = ω, .
,
(7.4)
Figure 7.2 shows the difference between the error functions assuming observation error
ω or process error only. It is visible on this …gure that in the case of process error,
the …rst summand (i = 1) is 0. Minimizing equation (7.4) is equivalent to a maximum
likelihood approach (Press et al. 1992) and since k is a constant it is also equivalent
to the traditional least-squares regression.
p
CVk2 + 1,
Notice that the expectations of and ω are not 1 but exp(k2 /2) =
k = ω, . This may seem strange on …rst view, but in fact allows the simple formulation
of the residuals as the difference of the log ’s. Had we E( ) = E(ω ) = 1, then we would
have to add k2 /2 to the residual (7.3) for it to have expectation 0 (Hilborn & Mangel
1997, personal communication with Ray Hilborn). The problem is purely technical:
one may prefer the lognormal variate to have expectation 1 or its log-transform to have
expectation 0. With real data one does not know which assumption is more reasonable.
Furthermore, since the k ’s are often not known very precisely, there is a risk of doing
more harm than good by adding the term k2 /2 to the residuals di,k . Therefore, one
usually …nds in the statistical literature the difference of the log ’s only (Ratkowsky
1983; Hilborn & Mangel 1997), and we will follow this safer approach.
If both types of errors are present simultaneously, y(ti ) also depends on the observation error in the data point s steps previously, Yi s . The statistical literature proposes
several solutions on how to treat this problem of ‘errors-in-variables’. Clutton-Brock
(1967) suggested to use weighted loss functions with the weights taking account of the
uncertainty in Yi s :
m
X
(y (ti ) Yi )2
CB1 =
wi
i=1+s
with
wi =
2
ω,i
+
2
ω,i
1
dy (ti )
dYi s
2
,
(7.5)
The last term in this equation, dy (ti )/dYi s , is the derivative of the predicted abundance
y(ti ) with respect to the initial condition y(ti s ) = Yi s . [Note that for y(ti ) = (Ni , Pi ),
the weight for state variable yj (ti ) is calculated by
2
0
ω,1,i s
∂Ni s yj (ti )
2
ωj,i = ω,j,i + (∂Ni s yj (ti ), ∂Pi s yj (ti ))
2
0
ω,
∂Pi s yj (ti )
2,i s
and adapted similarly for higher dimensional state variables]. The standard deviation
of the observation error, ω,i , must be known in advance (by multiple samples) and
104
Identifying predator-prey models (PhD Thesis)
C. Jost
independently of the process error. Here, we assume that the observation error has a
constant CVω known from replicate measurements. The standard deviation can thus be
approximated by ω,i = CVω Yi .
Another loss function, similar to the negative log-likelihood, was also introduced by
Clutton-Brock (1967),
CB2 =
m
X
0.5
i=i+s
(y (ti ) Yi )2
+ log(2wi )
wi
with wi de…ned as in equation (7.5).
Note that in error functions CB1 and CB2 the residual is no longer the difference
of the log ’s as suggested with the lognormal error type. However, the property of
the lognormal error, that the standard deviation is proportional to population size, is
preserved. We have also tested the lognormal versions of these equations, as proposed
by Clutton-Brock (1967) and by Carpenter et al. (1994), but these functions converged
very often to strange solutions, maximizing the dependence on the initial condition
(dy(ti )/dYi s ) rather than minimizing the residuals. They also converged much slower.
Using functions CB1 and CB2 thus simpli…es the regression task without losing much
generality with respect to the error type.
In sum, if the source of error in the data is assumed to be observation error only,
the function Xω2 must be used as a regression criterion. If it is thought that process
error only is present, the criterion X2 must be used. And if both errors are present
simultaneously, CB1 or CB2 can be used. In our study of identi…ability with arti…cial
data, we will consider all situations and we will assess empirically the discriminative
performance of all four error functions.
7.3.3
Model selection
The quality of adjustment of models to data is assessed with the familiar sum of squares
X 2 (7.4). This selection criterion is identical to the regression criterion when regression
has been done using Xω2 or X2 . The error functions CB1 and CB2 cannot be used directly
for model selection because the estimators use weightings that differ among models
(Carpenter et al. 1994). We based therefore model selection for all error functions on
the sum of squared residuals of log-transformed values (7.4).
To quantify the identi…ability of the models, we calculate the ratio Xc2 /Xf2 for each
time-series with Xc2 being the sum of squares after …tting the correct model and Xf2 the
sum of squares after …tting the false model (i.e., a ratio smaller than one indicates that
the correct model has been identi…ed). Plotting the cumulative distributions of these
ratios gives a visual representation of the selection performance of the different types of
error functions.
If the model is linear in the parameters, then the probability distribution of X 2 near
the optimal parameter values is the chi-square distribution with DF = 2m p degrees
of freedom (p being the number of parameters). A convenient way to calculate this
Jost, Arditi
Identifying predator-prey models I
probability is to use the incomplete Gamma function
`=
X 2 DF
,
2
2
105
(Press et al. 1992):
.
(7.6)
` is a measure of the improbability that this …t was obtained by chance. Although
this equation gives the correct improbability only for models that are linear in the
parameters, it is quite common to use it also with nonlinear models (Press et al. 1992).
However, in this case, ` is only a measure of this improbability. The actual values of
` obtained after …tting our models to the arti…cial data sets will give an indication
of the threshold value below which …ts to real data should be rejected because the
improbability that the …t was obtained by chance becomes too low.
7.3.4
Parameter estimation
Before trying to …t our six-parameter models, it should be veri…ed that the models are
well de…ned, in the sense that parameters are uniquely identi…able (Walter 1987). Based
on a Taylor-series expansion approach, this problem can be reduced to the algebraic
problem of solving six equations for six unknowns. It can be shown (C.2) that, in both
models, the six parameters are indeed uniquely identi…able if the state variables are
known with arbitrary precision and arbitrary resolution in time.
Since system stability is of much interest in ecosystems (return time after perturbations, persistence in stochastic environments), the local stability of the non-trivial
equilibrium will be used as an overall measure of the correct estimation of the entire
parameter set. Stability is measured by Re(), with being the dominant eigenvalue
of the community matrix (the Jacobian at the equilibrium point). This value has been
calculated analytically (to reduce numerical roundoff errors). We will plot the cumulative distribution of the ratios, |Re(e )/Re(c )|, with e the dominant eigenvalue for the
estimated parameters, and c the dominant eigenvalue for the correct parameters. The
steepness of this curve indicates the variation in the estimation of stability and, if the
curve passes through the point (1.0, 0.5), then there is no deviation from the expected
median.
The quality of the individual parameter estimates will be assessed by computing their
coefficients of variation from the …ts to each set of …ve replicated time-series. Averaging
these CV ’s over all parameter sets will give a general idea of the quality to be expected
with this type of …tting and data.
7.3.5
Algorithmic details
There does not exist much customizable software that allows …tting differential equations
to data. Fitting and visualization are usually separate steps in most software, which
further slows down the …tting process. Therefore we programmed the whole procedure
directly in C++ to create an application that allows immediate visual control of the
…tted model. This proved to be an indispensable tool to analyze large numbers of data
106
Identifying predator-prey models (PhD Thesis)
C. Jost
sets. (The software can be obtained from the …rst author upon request, and it requires
a Power Macintosh.)
In a …rst step, the time-series data were used to determine upper and lower bounds
of the parameters. These bounds were found by …rst computing a rough estimate of
each parameter. For r, a (), and h, this was done by an analogy with exponential
growth:
y˙ = ry ⇔ r =
log(y (ti+1 ))
ti+1
log(y (ti ))
.
ti
For example, a rough estimate of the maximal prey growth rate r was obtained by
calculating
log(Ni+1 ) log(Ni )
max
.
1im 1
ti+1 ti
The parameter K was roughly estimated as the maximal prey abundance. These estimates were then multiplied by some constants to get upper and lower bounds. The
constants were calibrated with the arti…cial data sets in such way that the intervals
contained the real parameters that had generated all these data sets. e was restricted
to the ecologically reasonable interval (0, 1).
In a second step, a genetic algorithm (GAlib 1.4.2 from http://lancet.mit.edu/ga/)
was used to search within these bounds, with population size 50, mutation rate 0.01,
crossover rate 0.1 and 400–600 generations. In this and the following step the solutions
of the ordinary differential equations needed to calculate Xk2 were simulated with the
adaptive stepsize …fth-order Runge-Kutta method odeint from Press et al. (1992).
In a third and …nal step, starting from the parameter values found by the genetic algorithm, the …tting was completed by using repeatedly a Levenberg-Marquardt method
and the downhill simplex method of Nelder and Mead combined with simulated annealing (routines mrqmin and amotsa from Press et al. 1992) until the …t could not be
improved any further. The Levenberg-Marquardt method requires at each step the simulation of a system of 20 (or more) coupled ordinary differential equations (see Appendix
7.A). The parameters were forced to remain within the calculated intervals during
the optimization (by blocking the parameter in the Levenberg-Marquardt method or
by penalizing the error function in the simplex algorithm, with a penalty that grows
exponentially with increasing distance from the bound).
The stopping criterion for both algorithms was determined dynamically from the
data set: let En be the error at step n (expressions Xk2 , CB1 or CB2 ), nj the estimate
of parameter j at step n (1 j p), and c = max 1im {Yi } CV m C (C being a
constant, set to 10 8 ). Then the algorithm was stopped if either
0
or
En
En+1
En
c
j j 0 max n j n+1 c
1ip n
Jost, Arditi
Identifying predator-prey models I
107
(Seber & Wild 1989). With simulated annealing, the error En might actually increase
at the beginning of the optimization process. Therefore, this algorithm was not stopped
if the …rst expression became negative.
In sum, both models were …tted with each of the four error functions to each of the
400 time-series with the following procedure: (1) Calculate upper and lower bounds for
the parameters, (2) run a genetic algorithm, (3) use alternatingly and repeatedly the
Levenberg-Marquardt method and the simulated annealing simplex algorithm with the
stopping criteria above until the error did not diminish any further.
7.4
Analysis and results
After …tting the correct models to all data sets with low and high noise, we computed
the mean ` over all realizations of ` (equation 7.6). When assuming observation error
only (Xω2 , calculated with a total CV = 0.2) this gave an approximate value ` 0.5 0.9
for the time-series with high noise. If we assume having process error in the data (X2 ,
calculated with a total CV = 0.2), then we obtain ` 0.03 0.5 for the time-series with
high noise. These values suggest that …ts to real data with `-values considerably smaller
than these should be rejected. Interestingly, the values of ` for the ratio-dependent data
were always much higher than those for the prey-dependent data. It seems that process
error of the same level (p in equations 7.2) adds on average less accumulated process
error in the ratio-dependent model than in the prey-dependent model.
Regarding the numerical efficiency of the algorithms, the Levenberg-Marquardt
search worked fast and efficiently close to the optimum (compared to the simplex algorithm), but it often failed when the starting values obtained with the genetic algorithm
were far from the optimum. In these cases, the simplex algorithm usually found the
basin of attraction much faster. The combination of both algorithms ensured almost
always the convergence to the optimum.
7.4.1
Model identi…cation
Figure 7.3 shows the cumulative distributions of the ratios Xc2 /Xf2 for all error functions
and noise levels. At low noise levels (CV = 0.05, p = 0.05) we see that error functions
Xω2 , X2 and CB2 led to less than 5% wrong identi…cations, while CB1 had about 15%
erroneous identi…cations. Therefore, we did not use the error function CB1 at the
higher noise levels (CV = 0.1, p = 0.1). At this higher noise, the error function Xω2
still had less than 5% wrong identi…cations, while the error function X2 had up to 10%
and function CB2 performed even worse. Therefore, we concluded that error functions
CB1 and CB2 are not useful for model selection, probably because the function that
is minimized and the model selection criterion are not the same. If we want 95%
con…dence in the identi…cation with error function X2 , then there can be at most 5%
wrong identi…cations. Estimating this from Figure 7.3(c) (by drawing a line at the 95%
level and projecting the intersection point with the distribution curve onto the x-axis
(point ), then taking the inverse of this value, 1/ ), we obtain the criterion that the
ratio of better …t by worse …t should be smaller than 0.95.
108
Identifying predator-prey models (PhD Thesis)
Prey-dependent data
Cumulative distribution
1
Ratio-dependent data
1
(a)
C. Jost
CV=0.05
= 0.05
(b)
0.5
0.5
0
0.0
0.5
1
1
0
0.0
1
(c)
1
0.5
1
(d)
CV=0.1
= 0.1
0.5
0.5
0
0.0
0.5
0.5
1
0
0.0
X2c/X2f
Figure 7.3: Quantile plots of …tting the models to the arti…cial data. — — —:
error function Xω2 ; ———: error function X2 ; - - - - -: error function CB1 ; – – –
–: error function CB2 . Xc2 is the error after …tting the correct model and Xf2 is
the error after …tting the wrong model. CV is the coefficient of variation of the
observation error and is the standard deviation of the process error. The dashed
straight lines in (c) show how the con…dence level of 95% for model selection with
error function X2 is calculated. See the text for more details.
7.4.2
Parameter estimation
Figure 7.4 shows the sensitivity of all four error functions to variations of one parameter
at a time (…xing the others at the values with which the time-series was created) for
a ratio-dependent data set with high noise (CV = 0.1, = 0.1). We see that they
are all quite unbiased with mostly symmetric error functions. Observation error Xω2
gave always the most narrow function, often asymmetric and the error increasing very
fast with the distance from the true parameter value [the steepest increase occurs if a
too efficient predator (high and e or low h, and r) drives the system to exinction].
This illustrates why we needed a genetic algorithm to …nd initial parameter estimates
within the basin of attraction of the optimal parameter values. But once this basin is
found, convergence with the simplex algorithm or the Levenberg-Marquardt method is
very fast. CB1 and CB2 show the ‡attest error functions, indicating slower convergence
rates of the optimization process. This same picture emerged with other data sets and
models.
Comparing directly the dominant eigenvalues of the non-trivial equilibrium with the
estimated and the correct parameters, strong differences between the error functions
Jost, Arditi
Identifying predator-prey models I
Superimposed error functions
r
1
1.2
1.4
K
1.6
1.8
600
800
1000 1200 1400
0.18
2
3
4
5
6
7
8
e
h
0.16
109
0.2
0.22
0.16
0.2
0.24
0.9
1
1.1 1.2
1.3
Parameter value
Figure 7.4: Four error functions vs. parameter values for a ratio-dependent timeseries with large noise. The curves where shifted vertically to put their minimum
all at the same height. Patterns are the same as in Figure 7.3. Actual parameters
are: r = 1.29, K = 900, = 4.7, h = 0.172, e = 0.207, = 1.01. They are
indicated by an arrow.
emerge. Figure 7.5 shows the cumulative distribution functions of the ratios of the
dominant eigenvalues for all estimated parameter sets, error functions and models. The
steepness of each curve is approximately the same, meaning that each error function
shows the same variation in the estimation of local stability, although there seems to be a
slightly smaller variation with ratio-dependent data. With respect to deviation from the
expected median, we see that error function Xω2 (…tting the whole trajectory) performed
overall best, followed by CB2 . Error function X2 always overestimated stability.
The coefficients of variation CV for each parameter, each model and each error
function (computed from the …ts to data with high noise only) are shown in Table 7.4.2.
The important point is that these values are generally high (15-90%), indicating that
even with 5 replicates, there remains much uncertainty in the estimated parameters.
The estimation of K and r were the most reliable, and all other parameters had CV ’s
above 40%.
7.5
Discussion
We addressed in this article the problem of model selection by …tting dynamic models
to predator-prey time-series that contain both observation and process error. Fitting
assuming observation error only (error function Xω2 ) allowed for the most reliable model
identi…cation with both noise levels. Figure 7.3 suggests that identi…cation should remain possible even with noise levels slightly higher than CV = 0.1 and p = 0.1 or with
some outliers in the data.
Identifying predator-prey models (PhD Thesis)
110
Prey-dependent data
1
C. Jost
Ratio-dependent data
1
(a)
(b)
Cumulative distribution
CV=0.05
= 0.05
0
1
0.6
0
1.4 Re( )/Re( )| 0.6
e
c
1.0
1
(c)
1.0
1.4
(d)
CV=0.1
= 0.1
1.8
0
0.6
1.0
1.4
0
Re( e)/Re( c)| 0.6
1.0
1.4
1.8
Figure 7.5: Quality of the estimated local stability: cumulative distributions of
the ratios of the real part of the estimated dominant eigenvalue Re(e ) by the real
part of the correct dominant eigenvalue Re(c ). Line patterns are the same as in
Figure 7.3.
Fitting assuming process error (function X2 ) leads to much less reliable identi…cation.
Identi…cation worked well with the low noise level. But for the higher noise level, the
ratio of the lower error by the larger error should be below 0.95 to have 95% con…dence
in the result. Higher noise levels or outliers in the data will further aggravate the
reliability of model selection. The error functions CB1 and CB2 that were supposed
Table 7.1: The mean coefficients of variation for the …tted parameters and the
mean standard deviations of the estimated dominant eigenvalues, . All values
are calculated from the results of …tting the data with high observation and process
error (CV = 0.1, p = 0.1) to the correct model. The …rst column indicates the
error function that was minimized.
prey-dep. data CVr CVK CVa CVh CVe CV Xω2
0.16 0.14 0.43 0.33 0.6
0.52 0.04
2
X
0.25 0.22 0.56 0.54 0.78 0.69 0.047
CB2
0.2
0.17 0.57 0.44 0.7
0.62 0.045
ratio-dep. data
Xω2
X2
CB2
CVr
0.31
0.35
0.38
CFK
0.28
0.14
0.24
CV
0.7
0.61
0.61
CVh
0.63
0.93
0.82
CVe
0.54
0.54
0.51
CV
0.59
0.5
0.44
0.036
0.027
0.031
Jost, Arditi
Identifying predator-prey models I
111
to take both observation and process error into account gave unreliable identi…cation
results, probably because the function that is minimized is not identical to the selection
criterion, as this is the case with with error functions Xk2 . Preference must be given to
functions Xk2 (or more general maximum likelihood approaches) that use for regression
and selection the same function.
Comparing our results with the work of Carpenter et al. (1994), we can notice that
model selection is more reliable for the continuous models studied here than for the
discrete models studied by these authors. However, Carpenter et al. …tted the prey
equation only (using predator data as input into this prey equation). Therefore, they
only selected for agreement with the prey dynamics, while we selected for agreement with
both prey and predator dynamics. These authors also state that manipulation of the
biological system is necessary to identify models. Our analyses show that it is needed
to have initial conditions far from the equilibrium state, in order to generate strong
dynamics of the system on its way back to the equilibrium. This can be accomplished
in natural lakes by stocking or in laboratory cultures (batch or chemostat) by using low
initial populations.
There is an interesting difference between the two models: the ratio-dependent timeseries were always identi…ed more reliably, in the sense that the difference between
the sum of squares X 2 after …tting both models was on average larger with ratiodependent data than with prey-dependent data (‘pushing’ the cumulative distribution
functions to the left in Figure 7.3). It seems that the ratio-dependent model is more
‡exible, adjusting itself more easily and with smaller residuals to given data. Carpenter
et al. (1994) had found the same difference (their Figures 3 B and E). This raises the
problem that prey-dependent time-series are more often (wrongly) identi…ed as being
ratio-dependent than the other way round.
A most interesting observation is that the error functions Xω2 and X2 made simultaneous wrong selections in only 2 of the 400 time-series. We can therefore conclude that
the most reliable model identi…cation can be obtained by …tting both error functions
and by accepting a selected model only if both functions give the same result. Unfortunately, observation error …t can become unreasonable in long time series because of the
accumulated process error that increasingly diverts the system from the deterministic
description.
Regarding the performance of the error functions with respect to individual parameter identi…cation, all error functions give parameter estimates close to the actual values
(Figure 7.4). However, Table 7.4.2 (calculated CV ’s from the replicated time-series)
shows that only parameters r and K are estimated with high precision, all others having large CV ’s. Assessing the quality of parameter estimates with the method of the
dominant eigenvalue (Figure 7.5) shows that, at the low noise level, this estimation is
quite reliable for all error functions but with the higher noise level, there is a large variation with a considerable deviation from the expected median (Figure 7.5). The error
functions X2 and CB1 overestimate local stability (the estimated dominant eigenvalue
is too negative).
There exist alternatives to the way we addressed the problem of model selection in
this article. In particular, one could take a versatile model that is either prey-dependent
112
Identifying predator-prey models (PhD Thesis)
C. Jost
or ratio-dependent, depending on a speci…c parameter value (e.g., using the models of
Hassell & Varley (1969) or DeAngelis et al. (1975)), and then directly estimate this
parameter. However, estimating it by …tting the whole model (as done in this paper)
will also result in a large uncertainty of the estimate, thus reducing the selective power of
this approach. Using Bayesian approaches to estimate posterior distribution functions of
this parameter are another possibility (Stow et al. 1995), but they require sophisticated
multidimensional integration techniques. The direct comparison performed in this paper
is more parsimonious and has the additional advantage of choosing between models that
can be incorporated into complex food webs (Arditi & Michalski 1995; Michalski &
Arditi 1995a).
In fact, there is no general statistical solution to the problem of …tting nonlinear
models to data that have both observation and process error. As stated by Pascual
& Kareiva (1996), the practical solution is often to …t as if there were only one type
of error in the data. If neither of the error types should be neglected we suggest to
use both types of …tting and base model selection on the joint result. This conclusion
will probably remain valid for systems that do not …t into the framework of this study
(e.g., systems that have only one state variable, unstable dynamics or more available
data points), but in these cases identi…ablility should again be veri…ed by a simulation
analysis similar to the one presented in this paper.
Our general conclusion is that, to address the question of identi…cation of dynamic
models, the scientist should …rst try to reduce observation and/or process error as much
as possible. If both errors remain important, then model selection is most reliable if both
observation error …t and process error …t select the same model. Parameter estimates
obtained by these methods are characterized by large coefficients of variation. The data
should also exhibit dynamics of much higher amplitude than the errors in the data.
This can be obtained either by low initial conditions in laboratory experiments or by
perturbation of natural systems.
Acknowledgements
We thank Jean Coursol and Brian Dennis for helpful and clarifying discussions on the
errors-in-variables problem. We also thank Eric Walter for useful comments on identi…ability and distinguishability. This research was supported by the Swiss National Science
Foundation (grant 31-43440.95 to RA) and by the French ‘Programme Environnement,
Vie et Société’ (CNRS).
Jost, Arditi
Identifying predator-prey models I
113
Appendix
7.A
Calculating the derivatives of the state variables with respect to the parameters
For the Levenberg-Marquart method as well as for calculating the error function CB1
and CB2 arises the problem of differentiating the solution of an ordinary differential
equation with respect to a parameter. We have applied the following method (see, e.g.,
Pavé 1994). Let y(t) be the solution of the differential equation
dy (t)
= f (y (t), a),
dt
y(t0 ) = y0
where f depends on a parameter a. We are looking for (t) :=
rule and changing the order of derivation, we obtain
dy (t)
.
da
Applying the chain
d dy (t)
d
=
(t)
da dt
dt
df (y, a) dy df (y, a)
=
+
dy da
da
df (y, a)
df (y, a)
=
(t ) +
dy
da
and, therefore, by solving the following coupled differential equations,
dy (t)
dt
d (t)
dt
y(t0 )
(t 0 )
= f (y (t), a)
df (y, a)
df (y, a)
(t) +
da
dy
= y0
= 0
=
we …nd the required derivative (t) by numerical integration. If a is the initial condition
y,a)
= 0 and (t0 ) = 1. This concept can easily be extrapolated to vector
y0 then df (da
valued y(t) (dimension m). In this case, dy/dt, dy/da, f , df /da, and d/da are vectors
of dimension m and df /dy is an m m matrix, the total derivative.
114
Identifying predator-prey models (PhD Thesis)
C. Jost
Chapter 8
From pattern to process:
identifying predator-prey models
from time-series data
Christian Jost, Roger Arditi
115
116
Identifying predator-prey models (PhD Thesis)
C. Jost
Abstract
Fitting time-series to nonlinear models is a technique of increasing importance in population ecology. In this paper we apply it to detect predator dependence in the predation
process by comparing two equally complex predator-prey models (one with and one
without predator dependence) to predator-prey time-series. Stochasticities in such data
come either from observation error or from process error or from both. We discuss how
these errors have to be taken into account in the …tting process and we develop eight
different model-selection criteria. Applying these to laboratory data of protozoan and
arthropod predator-prey systems shows that they have little predator dependence, with
one interesting exception. Field data are more ambiguous (either selection depends on
the particular criteria or no signi…cant differences can be detected) and we show that
both models …t reasonably well. We conclude that simple systems in homogeneous
environments show in general little predator dependence. More complex systems show
signi…cant predator dependence more often than simple ones, but the data are also often
inconclusive. Predictions for such systems based on simulation analyses should rely on
several models to reduce mathematical artefacts in these predictions.
Les techniques d’ajustement de modèles non linéaires à des séries temporelles prennent une importance croissante en écologie. Dans le présent travail, nous les appliquons
à la détection d’un éventuel effet de l’abondance des prédateurs sur le processus de
prédation en comparant deux modèles proie-prédateur d’égale complexité (l’un ave et,
l’autre sans prédateur-dépendance) à des séries temporelles d’abondances de proies et
de leurs prédateurs. La variabilité des données provient d’erreurs d’observation, d’erreurs de processus, ou des deux. Nous discutons la façon dont il faut tenir compte de
ces erreurs lors de l’ajustement de modèles, et nous développons huit critères différents
de sélection de modèle. L’ajustement des modèles à des données sur des systèmes proieprédateur de laboratoire (protozoaires et arthropodes) montre, à une exception près,
que l’abondance des prédateurs a peu d’in‡uence. Les données de terrain sont plus
ambiguës (soit la sélection dépend du critère de sélection retenu, soit la qualité de l’ajustement n’est pas signi…cativement différente pour les deux modèles), mais les deux
modèles s’ajustent correctement. Nous concluons que des systèmes simples en environnement homogène présentent en général peu de prédateur-dépendance. Des systèmes
plus complexes présentent plus souvent une prédateur-dépendance signi…cative que les
systèmes simples, mais les données ne permettent souvent pas de trancher. Il convient
donc de baser d’éventuelles prédictions fondées sur des simulations sur plusieurs modèles
a…n de réduire les risques d’artéfacts mathématiques dans ces prédictions.
Jost, Arditi
8.1
Identifying predator-prey models II
117
Introduction
The relation between predator-prey theory and real population time-series has been
the subject of many studies since the early publication of the Lotka-Volterra equations
or the Nicholson-Bailey model. The studied systems range from protozoan organisms
(Gause 1935; Luckinbill 1973; Veilleux 1979) over arthropod systems (Utida 1950; Huffaker 1958; Begon et al. 1996b) and microtine systems (Hanski & Korpimäki 1995) to
the whole plankton community of lakes (Scheffer 1998). Traditionally, model validation is done by comparing the data to the model either qualitatively (stable or cyclic
dynamic behaviour, length of cycles, amplitudes, etc.) or quantitatively (estimating parameters in the …eld, calibrating the model ‘by hand’ to obtain a good …t to the data).
Recent computer power combined with powerful global optimisation algorithms enable
researchers to …t rather complex mechanistic nonlinear models to time-series data. Such
…ts serve not only for model validation, but also for the detection of chaos (Turchin &
Taylor 1992; Dennis et al. 1997; Turchin & Hanski 1997), the testing of hypothesis
(Berryman 1996; Turchin & Ellner (unp.)) or model selection (Morrison et al. 1987;
Carpenter et al. 1994; Harrison 1995; Morris 1997). See Shea (1998) for a review of
general uses in population ecology.
In this paper we use model …tting as a criterion for model selection. The particular
functional form of a predator-prey model can have implications in …sheries management
and conservation biology (Yodzis 1994), on persistence of populations (Myerscough et al.
1996) or on spatial distributions of predators (van der Meer & Ens 1997). We address the
question of detecting predator dependence in the functional response (e.g., due to interference amongst searching predators). The functional response links prey and predator
dynamics in all models that follow the conservation of mass principle (Rosenzweig &
MacArthur 1963; Ginzburg 1998). In this large general framework, many different expressions for the functional response can be found in the literature (see May (1976b) and
Michalski et al. (1997) for inventories). The most widely used forms (Lotka-Volterra,
Holling type I, II and III) are functions of the prey abundance only (prey-dependent,
termed “laissez-faire” by Caughley (1976)) and do not depend on predator abundance.
Expressions that include predator dependence become usually more complex (Hassell &
Varley 1969; Beddington 1975; DeAngelis et al. 1975) which renders parameterization
or theoretical analysis more tedious. In practice this often results in the use of (simpler)
prey-dependent functional responses. The discussion on the importance of predator dependence has been revived by the introduction of the ratio-dependent concept (Arditi
& Ginzburg 1989) which offers a theoretical framework for modelling predator-prey
systems (but also food chains or whole food webs) with a functional response that inherently includes predator dependence while preserving the simplicity of the traditional
Holling-type functions. While prey-dependent predator-prey models rest essentially on
top-down mechanisms (Oksanen et al. 1981), ratio-dependent models can re‡ect both
bottom-up and top-down relations (Arditi & Ginzburg 1989; Poggiale et al. 1998).
The two views have mostly been tested by comparing equilibrium population abundances along a gradient of enrichment or by reanalyzing data of published functional
response experiments (see next section). Fitting models to time-series data of populations that are not in an equilibrium and applying goodness-of-…t as a criterion ap-
118
Identifying predator-prey models (PhD Thesis)
C. Jost
proaches the problem from a dynamic point of view. The way how models should be …t
to time-series depends on the source of errors in the data (Solow 1995; Hilborn & Mangel 1997). If the underlying process is stochastic and there is no error due to imperfect
sampling (observation error), then predictions are only possible for a limited time into
the future, e.g., to the next data point (1-step-ahead …tting). On the other side, if the
underlying process is deterministic and there is only observation error, then we can …t
the population trajectory (as determined by the model, its parameters and the initial
population size) over the whole length of the time-series (see Figure 8.1). We will adopt
Pascual & Kareiva’s (1996) terminology for these two types of …tting, calling the …rst
process-error-…t and the second observation-error-…t. In the literature one may
also …nd the terms s-step-ahead …tting (Ellner & Turchin 1995) for process-error-…t
and trajectory …tting for observation-error-…t.
A process-error-…t approach was used by Carpenter et al. (1994) for 7 years of
freshwater plankton data in two North American lakes (Paul Lake and Tuesday Lake),
one of them having been manipulated during the experiment (addition and removal
of …sh). These authors …tted alternative prey- and ratio-dependent discrete models
(difference equations) assuming the (non-manipulated) parameters to be the same over
the whole 7 years. They showed in a simulation study that model selection should
be possible with the data of the manipulated system, but the actual real data analysis
yielded little interpretable results. One problem in their study is that plankton dynamics
are more correctly modelled by a continuous system (large populations and overlapping
generations) and taking the time between measurements as the prediction time step is
an arbitrary choice. Furthermore, parameters might change from one year to the next.
Another study (Harrison 1995) used an observation-error-…t approach and compared
several continuous predator-prey models (differential equations) by …tting them to the
protozoan data of Luckinbill (1973). Working with laboratory data, the author assumed
stochastic in‡uences to be negligible compared to observation error. A major problem
in this study is that Luckinbill did not provide any information on measurement error
in his data and therefore criteria that take model complexity into account (e.g. Akaike’s
information criterion, see Hilborn & Mangel (1997)) were not used by Harrison. Unsurprisingly, the model with the largest number of parameters gave the best …t.
The two described methods of …tting, i.e., process-error-…t and observation error
…t, are statistically correct only if there is either no measurement error or no process
stochasticity respectively. Methods that account for both errors simultaneously exist,
but they require independent estimation of one of the errors or of their relative size
(Clutton-Brock 1967; Reilly & Patino-Leal 1981; Schnute & Richards 1995; Pascual
& Kareiva 1996). Since this information is often not available or hard to obtain, the
“practical decision usually involves choosing between the two …tting procedures” and
“the two assumptions are expected to provide two extremes in a range of likely parameter
estimates” (quoted from Pascual & Kareiva (1996), who make an extensive discussion
of statistical properties of the two types of …tting).
In this paper, we use the method of …tting alternative models to time-series data in
order to address the biological problem of detecting predator dependence in the functional response. This will be done by …tting a prey-dependent and a ratio-dependent
continuous predator-prey model with the same number of parameters (analogously to
Jost, Arditi
Identifying predator-prey models II
119
the treatment of Carpenter et al. (1994) with discrete time models) to a large number of real population time-series and applying goodness-of-…t as a criterion. These
time-series come from simple protozoan batch cultures, spatially more complex laboratory arthropod systems and complex lake plankton systems. Since such data contain
both observation and process errors, we will apply systematically both process-error-…t
and observation-error-…t, assuming that selection of the same model with both types
of …ts indicates a more reliable result. This reliability is further extended by not using
least squares only but also robust techniques (to detect artifacts caused by outliers in
the data) and by the use of bootstrapping to test whether the differences found in the
goodness-of-…t or the predictive power are signi…cant. Since all time-series analysed
here have the characteristics of a continuous system (large populations, overlapping
generations), we …t the differential equations and not their discretised analogues. We
will work with data found in the literature (including the data of the studies mentioned
above) and original plankton data from Lake Geneva that all give reasonable …ts with
both types of …tting. We will reanalyze Carpenter et al.’s data with continuous models
and by …tting each year separately. Since seasonal plankton dynamics are often explained by deterministic models of the type used here (Scheffer 1998), it will also be
interesting to see how such models perform when confronted with real data. Harrison’s
analysis will partially be redone with models of the same complexity only (thus permitting direct comparison) and by testing his observation-error-…t results against process
error …t results. Based on all these …ts we will also discuss some of the advantages and
disadvantages of either process-error-…t or observation-error-…t.
The time-series being used are characterised by a small size [10 - 50 data triplets
(time, prey, predator)] and by rather large measurement errors with coefficients of variation (CV ) of up to 50 %. For these reasons, a good …t of some model does not
necessarily mean that the used formalism describes the biological processes correctly.
This was demonstrated convincingly for the case of logistic population growth already
very early in the history of mathematical population biology, namely by Feller (1939).
Besides Verhulst’s logistic function, he also considered two other mathematical models
that are S-shaped and that have the same number of parameters and showed that the
two forms …t equally well to the real data that were then considered to be the “proof”
that the logistic model has the character of a physical law. This example illustrates
that obtaining a good …t of a model to data is not a proof of the biological correctness
of the chosen mathematical formalism. Or, in the words of Cale et al. (1989), ‘multiple
process con…gurations can produce the same pattern’.
These serious problems of model …tting can be addressed by following May’s (1989)
advice by ‘generating pseudo-data for imaginary worlds whose rules are known, and
then testing conventional methods for their efficiency in revealing these known rules’.
Following this advice and Carpenter et al.’s example, we tested the distinguishability
of the two models under consideration in a simulation study (see chapter 7). This
study showed that the tested models can reliably be identi…ed by goodness-of-…t from
predator-prey time-series (that reach an equilibrium after at least one large amplitude
oscillation) with 20 data triplets and moderate errors (CV of 10%).
120
8.2
Identifying predator-prey models (PhD Thesis)
C. Jost
The alternative models
Based on the principles of mass conservation and decomposition of the dynamics of a
population into birth and death processes we write the canonical form of a predator-prey
model as
dN
dt
dP
dt
= f (N )N
g(N, P )P
= eg (N, P )P
P,
(8.1)
where N and P are prey and predator abundances respectively, f is the prey growth
rate in the absence of a predator, the predator mortality rate in the absence of
prey and e the conversion efficiency. Predation is represented in these equations by
the functional response g (Solomon 1949), which in general depends on both prey and
predator abundances.
In order to …t model (8.1) to data, we have to formulate f and g explicitly. For the
recruitment function f , we use a standard logistic growth,
f (N ) = r(1
N
),
K
with intrinsic growth rate r and carrying capacity K . For the functional response two
models will be considered, a prey-dependent one and a ratio-dependent one:
aN
1 + ahN
g(N, P ) →
N/P
,
1 + hN/P
(8.2)
where a is the searching efficiency, h the handling time and an overall searching ef…ciency for all predators. The dynamics of the resulting predator-prey systems can
be, with both functional responses, stable coexistence, unstable coexistence (limit cycles) or extinction of the predator. The ratio-dependent model also offers extinction
of both prey and predator (Jost et al. 1999). These dynamics are also observed in
the time-series with which we shall compare the models by a goodness-of-…t criterion.
For this comparison to be possible, even in the absence of exact knowledge about data
quality (type of error, standard deviations), the models must have the same number of
parameters. As mentioned in the Introduction, there exist various models with various
degrees of predator dependence. The two above are taken as examples at the two extremes (8.2). We chose the Holling type II model because it is the most widely used
predator-independent (= prey-dependent) functional response. The ratio-dependent
model is the simplest predator-dependent functional response that has the same number of parameters as Holling’s model and that offers comparable dynamics. Alternative
predator-dependent models with comparable dynamics (Beddington 1975; DeAngelis
et al. 1975; Hassell & Varley 1969) have more parameters.
The ratio-dependent model was originally proposed as a simple hypothesis that accounts for predator dependence (Arditi & Ginzburg 1989). The fact that it was justi…ed
with empirical and phenomenological arguments has aroused some controversy (Abrams
1994; Ruxton & Gurney 1994; Akçakaya et al. 1995; Abrams 1997) but mechanisms
Jost, Arditi
Identifying predator-prey models II
121
leading to ratio dependence have now been demonstrated (Poggiale et al. 1998; Cosner
et al. in press). However, it should be noted that in highly complex systems, such
as lakes, there can be too many important processes at work to incorporate them all
in a mechanistically-derived model (spatial aggregation, defence mechanisms, refuge,
etc.). In such situations, we consider that phenomenological models whose predictions
correspond to global empirical patterns are reasonable options.
The two contrasting hypotheses have mainly been tested indirectly, by comparing
equilibrium properties in relation to enrichment, based on empirical evidence for positive correlation between trophic level abundances in freshwater lakes (McCauley et al.
1988; Arditi et al. 1991a; Mazumder & Lean 1994; Mazumder 1994; McCarthy et al.
1995). There is also experimental evidence that spatial heterogeneity induces predator
dependence in the functional response (Beddington et al. 1978; Arditi & Saïah 1992).
Analysis of direct measurements of the functional response at the behavioural level has
revealed many cases of predator dependence with some cases agreing with ratio dependence (Arditi & Akçakaya 1990). Predator dependence seems thus to be a common
occurrence in natural populations. Ratio dependence is just one particular way to include predator dependence, but it does so in a parsimonious way and allows for direct
comparison with prey-dependent types of the functional response. Therefore, a better
…t of the ratio-dependent model over the prey-dependent model cannot be interpreted
as a proof that this model is correct, but only that it better approximates the actual
occurrence of predator dependence.
We have avoided so far the term “density dependence”. There is some confusion in
the literature about the use of this term in the context of predator-prey models since
density can refer to prey or predator abundance. We agree with Ruxton and coworkers
who equate density dependence with predator dependence (Ruxton & Gurney 1992;
Ruxton 1995), because the functional response is intrinsically bound to the predator
and also because density dependence indicates in its usual sense a detrimental effect
of density on its own population’s growth rate. However, some authors refer to prey
density when using the term ‘density-dependent predation’. We consider this to be an
incorrect usage of the term, but to avoid any confusion, we will refrain from this term.
8.3
8.3.1
Materials and methods
Time-series
Two kinds of time-series data are analysed: data retrieved from the published literature
and unpublished original data of phyto- and zooplankton dynamics in Lake Geneva.
Only data with strong dynamics (sustained or damped oscillations) and allowing reasonable …ts with both process-error-…t and observation-error-…t are considered. Since
the difference between the two functional responses we compare is in the in‡uence of
predator abundance, we request that the range of variation of predator abundance over
time is such that the CV is greater than 50%. See Table 8.1 for a listing of all data
taken from the literature. The data were usually obtained by scanning the graphics and
extracting the data with the software DataThief (Macintosh). This process introduces
122
Identifying predator-prey models (PhD Thesis)
C. Jost
unavoidably some error, but this error is of minor importance compared to the …nal
residuals in the …ts. The data for Paul and Tuesday Lake (Carpenter et al. 1993) were
given directly in tabulated form within the publication, and some missing data points
were obtained from the authors. One data set of Luckinbill (1973) was also obtained directly from the author. Most of these data sets were originally given with no indication
on the observation error, or it was measured once and assumed to be similar for all data
points (Carpenter et al. 1994; Huffaker 1958). Therefore, the calculated likelihoods (see
below) cannot have an absolute meaning, and will only serve to reject …ts if the model
becomes too unlikely.
Table 8.1: Sources for data are: 1) Gause (1935), 2) Gause et al. (1936), 3)
Luckinbill (1973), 4) Veilleux (1979), 5) Flynn & Davidson (1993), 6) Wilhelm
(1993), 7) Huffaker (1958), 8) Huffaker et al. (1963), 9) Carpenter & Kitchell
(1994), 10) CIPEL reports (1986-1993). Numbers or letters separated by comma
refer to further data sets with the same basic name, e.g. gause1,-3 refer to data
sets gause1 and gause3. Usually these numbers refer to the …gures or tables within
the cited publication, except for plankton data where they refer to the year the
data have been collected.
Name of data set
source prey
predator
type of data
Paramecium
Didinium
gause1,-3,-4
1
batch culture
caudatum
nasutum
(#individuals)
Aleuroglyphus Cheyletus
gauset2a,-c,-d,-e,-f
2
batch culture
agilis
eruditus
(#individuals)
Paramecium
Didinium
luckin2a,-2b,-3,
3
batch culture
aurelia
nasutum
-4a,-4b,-5
(#individuals)
Paramecium
Didinium
veill8, -10
4
batch culture
aurelia
nasutum
(#individuals)
Isochrysis
Oxyrrhis
‡ynn1b,-1c,-2b,-2c
5
batch culture
galbana
marina
(#cells)
Escherichia
Tetrahymena
wilh4.2,-4.4,-5.27,
6
batch culture
coli
thermophila
-5.28,-5.29,-5.30
(biovolume)
Eotetranychus Typhlodromus laboratory
huff11,-12, . . . ,-18
7
sexmaculatus
occidentalis
(#individuals)
Eotetranychus Typhlodromus laboratory
huff63-3,-63-4
8
sexmaculatus
occidentalis
(#individuals)
paul84,-85, . . . ,-90,
9
edible
zooplankton
whole lake data
tues84,-85, . . . ,-90
phytoplankton
(biomass)
edPhy86,-87, . . . ,-93
10
edible
herbivorous
whole lake data
phytoplankton zooplankton
(biomass)
totPhy86,-87, . . . ,-93
10
total
herbivorous
whole lake data
phytoplankton zooplankton
(biomass)
The data on Lake Geneva were collected as part of the lake monitoring program of the
International Commission for Protection of Lake Geneva Against Pollution (CIPEL).
The sampling methods are described in annual reports, e.g. CIPEL (1995). A short
Jost, Arditi
Identifying predator-prey models II
123
description can also be found in Gawler et al. (1988). Phytoplankton was sampled with
a Pelletier bell-shaped integrating sampler from 0 to 10 m water depth. Zooplankton
was sampled by vertical tows from a depth of 50 m with coupled nets. Plankton biomass
was calculated from abundance and estimated biovolume. Phytoplankton with length
< 50 m and biovolume < 104 m3 was considered edible phytoplankton. Herbivorous
zooplankton was identi…ed by species and age class: cladocerans (mainly Daphnia and
Bosmina), calanoides (Eudiaptomus) and cyclopoids for the age classes nauplii to copepodites stage 3 (higher age classes were considered carnivorous). The samples were taken
at the station SHL2, at the centre of the lake, midway between Evian and Lausanne
(lake depth 309 m). Plankton was usually sampled twice a month. The observation
error was not measured, but the collecting scientists estimate the CV to be in the range
of 10-20 % for an individual sample as in Carpenter et al. (1994). However, there are
important heterogeneities in the lake, and the CV between several samples in the same
area at the same time can be much larger (N. Angeli, pers. comm.). For this reason, we
use in the …tting a prudent CV of 50 %, which is considered realistic for zooplankton
but somewhat pessimistic for phytoplankton.
8.3.2
Error functions
As explained in the Introduction, we consider two types of errors, measurement error
(imprecise sampling) and dynamic noise (due to environmental stochasticities, differences between the biological process and its mathematical description and, to a lesser
extent in our data, demographic stochasticity). Figure 8.1 illustrates the resulting difference between observation-error-…t and process-error-…t. The prediction horizon with
process-error-…t is traditionally one time step ahead (e.g., Carpenter et al. 1994, Dennis
et al. 1995). For the reasons detailed in Ellner & Turchin (1995), we predict the dynamics s time steps ahead, with s chosen in such way that the autocorrelation with the
predictor drops under 0.5. The main argument for this choice is to detect the nonlinear
dynamics in the time-series. On smaller prediction horizons, standard statistical linear
regression models might be more powerful than our mechanistic nonlinear models. See
Ellner & Turchin (1995) for a more detailed justi…cation.
We will assume that the (observation or process) error in the densities is of a multiplicative type (lognormal), with a constant CV (stationarity). This type of distribution
is typical for population data in general (Dennis et al. 1997; Hilborn & Mangel 1997).
It was used and justi…ed for plankton in particular by Carpenter et al. (1994). For protozoan and arthropod data also, this choice seems reasonable (Wilhelm 1993; Huffaker
1958). In general, the true nature of the error can be intermediate between normal and
lognormal. This was assumed in the regression analysis of Harrison (1995) performed
with Luckinbill’s data, but this approach requires another parameter that Harrison determined empirically. For reasons of parsimony, we refrained from such an approach,
and the lognormal error type seems on the whole better suited to the kind of data we
analysed. Cases in which model selection might be biased by using lognormal error
are discussed speci…cally. Uncertainties in the time measurements are considered to be
small in comparison with the errors in abundances and are neglected in this study.
To formalise these ideas, consider a simple differential equation dy/dt = f (y, ) with
124
Identifying predator-prey models (PhD Thesis)
N(t)
C. Jost
Observation error fit
(a)
t
N(t)
Process error fit
(b)
t
Figure 8.1: Fitting a differential equation to a time-series. In (a) all error is
assumed to be measurement error and the whole trajectory is …tted at once
(observation-error-…t), while in (b) all error is assumed to be process error and each
data point serves as an initial condition to predict the next data point (processerror-…t).
time-series data (ti , Yi )1im , where ti is the time at which the population y is observed
to have density Yi , m is the number of data points [for predator-prey models y has to be
replaced by the couple (N, P )] and is a vector of q parameters. Let yˆi be the unknown
real population density at time ti , and y(ti ) = y(ti , ) the deterministic solution of the
differential equation at time ti with parameters . If there is only observation error
in the data, then the whole trajectory is …t and there is only one initial condition,
y(t0 ) = yˆ0 , that is treated like a free parameter. If there is only process error then the
initial condition is different for each predicted data point and de…ned as the data point
s steps previously, y(ti s ) = Yi s , s 1. With these notations and with ti = ti ti s ,
the process can be written as
observation error only
yˆi = y(ti , )
Yi = yˆi ω
y(t0 ) = yˆ0
process error only
yˆi = y(ti , ) (ti , f, )
Yi = yˆi
y(ti s ) = Yi s
Process error and observation error ω are considered to be lognormal random variates
with constant coefficients of variation CV and CVω respectively. However, note that
Jost, Arditi
Identifying predator-prey models II
125
the process error depends in general on the time interval between measurements and
on the underlying process f with parameters .
The log-transformed data
p are assumed to be Gaussian with expectation log(yˆi ) and
standard deviation k (= log(CVk2 + 1) ), k = , ω (Carpenter et al. 1994). We de…ne
the residual as
di = log(Yi )
log(y(ti )).
The …tting procedure minimises the sum of squared residuals weighted by the variances,
Xk2
m
X
d2i
,
= min
2
i=1+s k
k = , ω
(Xk2 )
(with s = 0 for k = ω). This least squares regression with the described statistical
model is equivalent to maximum likelihood (Press et al. 1992).
While the proportionality between population size and standard deviation of measurement or process error is generally accepted, there may be considerable doubt about
the normality of the log-transformed values. Regression with criterion Xk2 is very sensitive to outliers (data points that are farther away from their real value than would be
expected with a normal distribution) with respect to parameter estimation and model
selection (Linhard & Zucchini 1986). We therefore used as a second criterion the Laplacian (or double exponential) error function,
XkL = min
m
X
|di |
,
k
i=1+s
k = , ω
(XkL )
which is more robust against such outliers (Press et al. 1992).
8.3.3
Simulation study, numerical methods and bootstrapping
The power of error function Xk2 for selecting the correct model from time-series data
has been tested in a simulation study (see chapter 7): we parameterized the two models
(8.1 and 8.2) randomly to create deterministic dynamics that resemble the observed
dynamics in a lake (reaching a stable equilibrium after one or two large amplitude
oscillations) and simulated stochastic data (20 data points) containing both process error
and observation error (both of a multiplicative type and with CV = 0.1, comparable
to the ones found in lakes). Both models were …tted to these arti…cial time-series with
error functions Xk2 in a 3-step-procedure: (a) computing upper and lower limits of the
parameters from the time-series data, (b) …nding starting values of the parameters with a
genetic algorithm and (c) computing the optimal parameters with the standard simplex
method coupled with simulated annealing (Press et al. 1992). In this regression scheme
we use only the available time-series data to estimate the parameters and with them
the discrepancy between model and data. Only logical constraints such as positivity of
parameters are applied. This simulation study showed that either observation-error-…t
or process error …t alone detect the correct model with more than 95% con…dence if
we require that the goodness-of-…t for each model (as estimated by Xω2 or X2 ) differ by
126
Identifying predator-prey models (PhD Thesis)
C. Jost
more than 5%. More importantly, in only 1% of all …ts was the wrong model identi…ed
with both criteria simultaneously.
All real time-series in this study are …tted with error functions Xk2 and XkL by
the same 3-step-procedure that was used in the simulation study, with process-error…t and observation-error-…t. In a further analysis we estimate the expectation E(Xk2 )
by residual bootstrapping as described in Efron & Tibshirani (1993). Based on the
bootstrap …ts we calculate the improved estimate of prediction error IEP E (Efron &
Tibshirani 1993) to compare the predictive power of the two models. See the Appendix
for a detailed technical description of these methods. As a by-product, the bootstrapping
permits to estimate for each time-series the variation in the estimated parameters.
8.3.4
Model comparisons
Model selection will be based on the goodness-of-…t according to criteria X 2 , X L or
E(X 2 ) and the prediction error criterion IEP E detailed above. The likelihood ` of
the model (given the data) can be calculated approximatively with X 2 and the degrees
of freedom DF [number of predicted data points minus number of parameters, i.e.,
DF = 2(m s) q] by the formula
2
X DF
`=
,
,
(8.3)
2
2
where is the incomplete gamma function (Press et al. 1992). We could not derive such
a function for the likelihood in case of Laplacian error. Therefore, the same function is
used to get at least an order of magnitude of the likelihood.
Note that equation (8.3) is absolutely correct for linear models only (Press et al.
1992). To know what likelihood levels should be expected with our nonlinear models
we have to rely on the likelihoods computed in the simulation analysis (chapter 7)
that ranged from 0.01 to 0.9 (that is, likelihood ` after …tting the model that was
originally used to simulate the arti…cial data). Assuming a larger discrepancy between
the theoretical model and the real process we accept all …ts with a likelihood above
0.001 (calculated with a ‘prudent’ CV of 0.5), otherwise model selection is considered
nonsigni…cant. While X 2 and X L must differ by at least 5% between the two models to
be considered signi…cantly different (see previous section), E(X 2 ) and IEP E (obtained
by bootstrapping) are compared directly with a standard t-test ( = 0.05) and model
selection is accepted as signi…cant if the likelihood of E(X 2 ) is also above the limit 0.001.
The discussion will be based solely on the signi…cant …ts that are indicated in Table 8.4
by a † or ‡ . The ‡ indicates strongly signi…cant …ts (` > 0.001 with CV = 0.1).
The comparison between the qualitative dynamic behaviour of the data and the
(winning) …tted model is used as a second criterion for the adequacy of the model.
These qualitative dynamics are classi…ed into strongly stable equilibria (ss) when the
trajectory converges to the equilibrium after at most one oscillation, stable equilibria
(st) if there is more than one oscillation before stabilisation, limit cycles (l), extinction
of the predator only (pe) and extinction of both populations (e). See Figure 8.2 for an
illustration of the different types.
Jost, Arditi
Identifying predator-prey models II
127
Figure 8.2: The different dynamic behaviours that are distinguished in the data
and in the …tted models. Each graph shows the phase space with isoclines of prey
(isoN ) and predator (isoP ) and an example trajectory. pe designates extinction of
the predator only, ss strongly stable systems, st stable ones, l limit cycles, and e
extinction of both populations (only possible in the ratio-dependent model).
128
8.4
Identifying predator-prey models (PhD Thesis)
C. Jost
Results
The detailed …tting results (model selection per type of …tting and per criterion) with
additional information on each time-series (size and apparent dynamics) are summarised
in Table 8.4 (see legend for more details). To facilitate their interpretation the essential
model selection results are condensed in Table 8.2 applying the following rules: for each
time-series a “winning” model is selected if all signi…cant …ts with both types of …tting
and with the four different criteria identify the same model, otherwise the time-series is
marked as ambiguous. For the lake data the same procedure is also applied separately
for both types of …tting. Furthermore, we indicate if the qualitative dynamics (Figure
8.2) of the signi…cant …ts correspond to the apparent dynamics of the timeseries. Since
extinction is not possible in the prey-dependent model we assume correct detection of
the qualitative behaviour if the …t shows limit cycle behaviour (the closest to extinction
that this model can produce).
Table 8.2: Summary of model selections: the …rst part of the table lists for each
group of time-series the number of times each model was selected by both …tting
types, with the number of correct qualitative dynamics in parentheses. The number of ambiguous model selections are listed in the last column. The second part of
the table treats process-error-…t (PEF) and observation-error-…t (OEF) separately
for lakes.
data
prey-dep.
ratio-dep.
ambiguous
Gause
6 (3)
1 (1)
1
Luckinbill, Veilleux
6 (1)
0
2
Flynn and Davidson
0
4 (4)
0
Wilhelm
2 (0)
2 (2)
2
Huffaker
6 (5)
0
4
Paul Lake
2 (2)
2 (0)
3
Tuesday Lake
4 (2)
0
3
Lake Geneva (edible)
2 (1)
4 (2)
2
Lake Geneva (total)
5 (2)
1 (0)
2
PEF OEF PEF OEF PEF OEF
Paul Lake
2 (2) 3 (3) 2 (0) 2 (0)
3
2
Tuesday Lake
2 (1) 6 (3) 1 (1)
0
4
1
Lake Geneva (edible) 4 (3) 2 (1) 1 (0) 6 (3)
3
0
Lake Geneva (total)
2 (2) 6 (2) 3 (1) 1 (0)
3
1
8.4.1
Protozoan data (simple batch cultures)
The protozoan data of Gause, Luckinbill, Flynn & Davidson, and Veilleux were the easiest to …t and they gave the most signi…cant results. Of Gause’s eight data sets, six are
prey-dependent, one ratio-dependent and one nonsigni…cant. The latter data set has
less than 10 data triplets (time, prey, predator); it may simply be too short for a reliable
model identi…cation. In most cases the dynamic behaviour is also correctly detected by
Jost, Arditi
Identifying predator-prey models II
129
the winning model. We can therefore conclude that Gause’s data do not indicate any
signi…cant predator-dependence in the functional response. This is con…rmed by Luckinbill’s and Veilleux’s data (both with similar organisms), in which six prey-dependent
time-series are identi…ed and two are ambiguous. However, the observed limit cycle
was only detected once by process-error-…t. The regression resulted mostly in dynamics with a stable equilibrium. One data set of Veilleux …ts signi…cantly better to the
ratio-dependent model with observation-error-…t: these data show fast convergence to
a sustained stable limit cycle with minima far above zero, a pattern that cannot be
produced with the prey-dependent model. However, the ratio-dependent …t does not
resemble the data very much either. We think that this limit cycle should rather be
explained by delayed effects, and the signi…cant ratio-dependent …t is also due to the
lognormal error structure that gives more importance to points closer to zero. When
repeating the regression assuming a Gaussian error the …t becomes indeed signi…cantly
prey-dependent.
An interesting exception to this prey-dependent predominance in protozoan systems
are the data of Flynn & Davidson (1993) which are all strongly signi…cantly ratiodependent for all criteria and both types of …tting. These data are those of static
batch cultures without aeration, stirred every day before sampling. It is therefore likely
that heterogeneities developed between stirrings. The data sets ‘‡ynn1b’, ‘‡ynn2b’ and
‘‡ynn2c’ are shortened from their Figures 1b, 2b and 2c respectively because there was
an obvious change of parameters during the experiment, detectable by an abrupt change
in dynamics and described by the authors as the onset of strong cannibalism amongst
predators. The data sets consist of the data before the change.
The data of Wilhelm (1993) are the least conclusive: two data sets are prey-dependent,
another two are ratio-dependent and two are ambiguous. The pre-equilibrium dynamics
in the time-series (always one large amplitude cycle followed by a long time of stable
coexistence) are probably too short for a reliable model identi…cation.
The variation in the estimated parameters is rather high (medians of the CV ’s
ranging from 0.16 to 0.78) but they are smaller than with the other time-series (see
below). Observation-error-…t gives slightly smaller CV ’s than process-error-…t.
8.4.2
Arthropod data (spatially complex laboratory systems)
Huffaker’s data …t rather badly to both models (low likelihoods), but the few signi…cant results are always prey-dependent observation error …t. Process-error-…t con…rms
this result with the exception of two signi…cant ratio-dependent …ts with Laplacian error
(equation XkL ). However, these two …ts are qualitatively wrong (stable dynamics instead
of limit cycles or extinction). The data themselves are always unstable with oscillations
and coming close to extinction, although the populations seem very often to recover
shortly before termination of the experiment. These dynamics are correctly reproduced
with both models and observation-error-…t. Interestingly the prey-dependent model retains these unstable dynamics with s-step-ahead …tting, while the ratio-dependent model
mostly converges to a stable system. This is a further indication that the prey-dependent
model is closer to the real dynamics. The medians of the CV ’s of the estimated para-
130
Identifying predator-prey models (PhD Thesis)
C. Jost
meters range from 0.09 to 1.91 (again with smaller ones in observation-error-…t).
8.4.3
Plankton data (complex lake plankton systems)
In general these data are very noisy and it is not clear if there are well de…ned dynamic
patterns or only noise. However, the algorithms seemed to …nd dynamic patterns in
some cases.
With observation-error-…t, Paul Lake gives …ve times a signi…cant …t, of which three
are prey-dependent. Process-error-…t gives four signi…cant …ts, also with two preydependent ones. In three cases, the two types of …tting are contradictory. Carpenter
et al. (1994) had not found any signi…cant result for this lake with discrete predator-prey
models and process-error-…t (one-step-ahead …tting), which agrees with our …nding. The
results are rather different in manipulated Tuesday Lake. Observation-error-…t selects
the prey-dependent model in six of the seven time-series, with one ambiguity. Processerror-…t is less conclusive, with two prey-dependent time-series, one ratio-dependent,
three nonsigni…cant and one ambiguity. Interestingly, Carpenter et al. found for this
lake a good …t to the ratio-dependent model. We do not know if this discrepancy comes
from the fact that we look at seasonal dynamics or because we use continuous models.
For Lake Geneva, we the phytoplankton either with edible algae only or with total
algal biomass. Observation-error-…t gives the clearest trends: ratio-dependent (6 vs.
2) for edible phytoplankton, prey-dependent (6 vs. 1) for total phytoplankton. With
process-error-…t the trends are less distinct but seem to contradict the previous results:
prey-dependent (4 vs. 1) with edible phytoplankton, ratio-dependent (3 vs. 2) with
total phytoplankton. Ambiguities between results from the two types of …tting appear
in four cases.
Regarding qualitative behaviour, observation-error-…t often results in transient dynamics without reaching an equilibrium state or the dynamics are in some cases true
limit cycles, while process-error-…t nearly always results in stable systems. Especially
the ratio-dependent model shows often strongly stable dynamics (a pattern termed “limited predation” by Arditi & Ginzburg (1989), see Figure 8.2). The medians of the CV ’s
of the estimated parameters cover a wide range from 0.4 to 2.4 (again with smaller ones
in observation-error-…t).
8.5
Discussion
We have compared predator-prey time-series (that contain both process and observation
error) to two alternative predator-prey models with the objective to detect predator
dependence in the functional response. The goodness-of-…t was estimated by …tting the
models to the time-series assuming that there is either observation error only or process
error only, an assumption made necessary by the insufficient quantitative information
about the actual errors in the data. Looking at the proportion of ambiguous model
selection due to a small difference between the two goodness-of-…t values (marked with
a in Table 8.4), we see that observation-error-…t was only half as much ambiguous as
Jost, Arditi
Identifying predator-prey models II
131
process-error-…t (in all analysed systems). It seems therefore that observation-error-…t
is a more efficient tool than process-error-…t to select models (this difference was also
noted by Harrison 1995). However, since the data contain both types of error, the most
reliable selection is possible if both types of …tting are used and if their selection results
are not contradictory. Actually, model selection is less limited by the type of …tting
than by the dynamic variation in relation to errors in the time-series. When there
are strong dynamics over the whole length of the time-series (Gause’s, Luckinbill’s,
Veilleux’s, Flynn and Davidson’s and Huffaker’s data) then both types detect the same
winning model and few ambiguous results due to small differences in goodness-of-…t are
obtained. Only if there is little dynamic variation (one large initial oscillation as in
Wilhelm’s data) or if the dynamics are hidden behind strong observation errors as in
lakes such ambiguous results become more frequent.
A number of technical problems with both types of …tting are worth discussing.
The strongest limitation for observation-error-…t is that it works only for relatively
short time-series because process error accumulates with time even in well controlled
laboratory systems (Harrison 1995). Process-error-…t is questionable with respect to
the choice of s (time steps ahead prediction, see discussion in Ellner & Turchin 1995).
Furthermore, when predicting s steps ahead, one implicitly also predicts s 1, s 2, . . . ,
1 steps ahead, and there is currently no statistical solution as to how this information
should be properly incorporated into the …tting process. For future research in this
direction, we need more quantitative information about observation or process error in
the time-series data. This would permit to select between models of different complexity
(by using information criteria such as Akaike’s information criterion, see Hilborn &
Mangel 1997) and the use of regression techniques that account for both observation and
process error (Reilly & Patino-Leal 1981; Schnute & Richards 1995; Ellner & Turchin
1995). However, these methods are computationally very expensive especially for …tting
continuous systems. Bootstrapping one single data set can take a full week of computing
time on a couple of workstations (Peter Turchin, personal communication). For such
reasons simpler regressions as done in this paper will remain a useful tool in ecology,
especially since extensive simulations with arti…cial data (May 1989) are often the only
possibility to test whether the available data can answer the question at hand.
The main subject of this study, detecting predator dependence in the functional
response, yields interesting and unexpected results. The most signi…cant results are obtained with the protozoan data. Most systems are either closer to prey dependence or
the samples are too small to detect reliably predator dependence. However, there is one
predator-prey system with four time-series (Flynn & Davidson 1993) that shows significant predator dependence. The predators in this system can show strong cannibalism
at low prey-densities (pers. comm. with the authors). Although such cannibalism was
not observed in the analysed data, this suggests that the predators are capable of strong
interference when they encounter one another. Since these batch cultures were stirred
only once every 12 h, it is possible that heterogeneities developed between stirrings.
Both factors have been shown to lead to predator dependence. To our knowledge this is
the …rst example of a protozoan system with monospeci…c prey and predator that exhibits so strong predator dependence that the ratio-dependent model …ts better. This
example demonstrates that the biology of the system must dictate the model being
132
Identifying predator-prey models (PhD Thesis)
C. Jost
built, and that traits like potential cannibalism can indicate that a model with predator
dependence is more appropriate.
Luckinbill and Veilleux obtained in their experiments several population cycles before
extinction by reducing the predator attack rate (thickening of the medium with methyl
cellulose) and by reducing prey carrying capacity and growth rate [using poorer nutrition (half strength cerophyl mixture) for the prey in time-series ‘luckin5’]. Qualitatively,
these parameter changes can stabilise both the prey-dependent and the ratio-dependent
model. Only by a direct quantitative comparison of the data with the two models can
we exclude predator dependence to be an important trait in these systems. The stable
limit cycles obtained by Veilleux (1979) (with a re…nement of Luckinbill’s technique)
differ qualitatively from both models, indicating that either a model with intermediate
predator dependence or other mechanisms such as delayed effects are important. Harrison (1995) obtained drastically improved …ts to Luckinbill’s data upon incorporating
these two traits (predator mutual interference and a delayed numerical response in the
form of nutrient storage).
It appears that simple homogeneous and monospeci…c predator-prey systems are
often better described by a prey-dependent model. Kaunzinger & Morin (1998) studied
a three-level protozoan food chain (bacteria - Colpidium striatum - Didinium nasutum)
and demonstrated that enrichment changes equilibria and system stability in a way that
agrees better with prey-dependent theory (Rosenzweig 1971; Oksanen et al. 1981). Bohannan & Lenski (1997) compared dynamics and equilibria of a bacteria-bacteriophage
system (in a chemostat setup, with two different nutrient in‡ow concentrations) qualitatively with a complex prey-dependent model (dividing bacteria into susceptible and
resistant strains) and with a simple aggregated ratio-dependent model (only one state
variable to describe bacteria). They found that the qualitative results (destabilization
with increasing nutrient in‡ow, both bacteria and bacteriophage equilibria increase with
the higher nutrient in‡ow) were better predicted by the prey-dependent model. However, the comparison is statistically unsatisfactory, since the number of parameters were
not taken into account (11 in the prey-dependent model and 7 in the ratio-dependent
model).
There also seems to be no predator dependence in Huffaker’s arthropod data. In
most cases the prey-dependent model …ts better, and in the two cases with process-error…t where the ratio-dependent model …ts better this …t is qualitatively wrong (stable
equilibrium instead of a limit cycle, see Table 8.4). Two other aspects are important
for the …ts to these data: (1) quantitatively the models …t rather badly to the data, the
experimental systems showing larger variation than can be reproduced by our simple
models and (2) observation-error-…t gives with both models qualitatively correct …ts
(see Figure 8.3). The …rst point might be explained by Huffaker’s experimental setup,
food for prey being dispersed in a 2- or 3-dimensional structure and the prey colonising
this food in a fairly heterogeneous manner. Such a laboratory system is structurally
more complex than the protozoan batch cultures of the previous paragraph. The second
point indicates that the used models can nevertheless be used for qualitative analysis;
only quantitative conclusions should be interpreted with care.
The …ts to phyto- and zooplankton data are the most difficult to interpret. The
easiest conclusion would be that either the data are too noisy for this kind of model
Jost, Arditi
Identifying predator-prey models II
5
2
10
4
15
6
8
133
20
10
12
Figure 8.3: Two examples illustrating (a) observation-error-…t with a prey-dependent model to Huffaker’s data and (b) process-error-…t (s = 4) with a ratiodependent model to Flynn and Davidson’s data.
identi…cation or that both models are too simple for lake dynamics. The …rst interpretation is supported by the qualitative nature of the process-error-…ts (mostly stable
or strongly stable systems) that might mean that the best prediction is not obtained
by dynamic nonlinear modelling but rather by simply using some mean abundance of
prey and predator. Despite these reservations, many signi…cant model identi…cations
were obtained with observation-error-…t, showing that long term dynamic patterns are
present. These signi…cant …ts are of both types (prey-dependent and ratio-dependent)
with tendencies for some lakes: Tuesday lake being mostly prey-dependent, the system with edible phytoplankton in Lake Geneva rather ratio-dependent and the system
with inedible algae more prey-dependent. However, these tendencies are not sufficiently
clear to give recommendations as to which model might be more appropriate. Brett &
Goldman (1997) argued that phytoplankton displays strong bottom-up in‡uence while
zooplankton is more sensitive to top-down control. The phytoplankton-zooplankton interaction itself (that is studied in this paper) is subject to both forces, which might also
explain the ambiguity in model identi…cation.
We can conclude that these heterogeneous systems with multispecies prey and predator levels show both types of functional responses or intermediate types. Or, in the words
134
Identifying predator-prey models (PhD Thesis)
C. Jost
of Yodzis (1994) who studied relations between predator dependence in the functional
response and …sheries management, “it remains frustratingly difficult to say just which
functional form is the appropriate one for a given population”. As a consequence, we
should base population management decisions on the predictions of several competing
models, building up con…dence in each model by constant comparison of its predictions
with actual observations. Decision making then results from the predictions of all these
models and on the current con…dence level in them (similarly to weather forecast). For
lake management in particular, other models could use sigmoid functional responses
since alternative prey exist [the ‘inedible’ algae can be consumed to some extent (Davidowicz et al. 1988; Gliwicz 1990; Bernardi & Giussani 1990)]. Simple linear forms of
the functional response are another reasonable choice in the context of process-error…t and short term prediction (Carpenter et al. 1994). A very good example of this
multi-model approach is given by Sherratt et al. (1997), who analyse four completely
different models (reaction-diffusion equations, coupled map lattices, deterministic cellular automata and integrodifference equations) to study invasion patterns in space. Since
a common feature emerges from all models, it can be regarded as a highly likely real
feature. Working with at least two models can help identify model artifacts and direct
further research.
With respect to detection of predator dependence, one might also suggest to consider an intermediate model that can be predator-dependent or predator-independent
depending on the value of a speci…c parameter (as done in Arditi & Akçakaya (1990) for
functional response data) and then directly estimate this parameter. However, looking
at the uncertainty of the estimated parameters in this study makes it unlikely to obtain
more information about predator dependence by this intermediate approach. Bayesian
approaches to estimate posterior distribution functions of this parameter were also proposed (Stow et al. 1995), but they require sophisticated multidimensional integration
techniques and little is known about the robustness of these methods when confronted
with ecological data of the type used here.
One basic support for the Holling type II function comes from an analogy with
the Michaelis-Menten enzyme kinetics and Monod’s work on bacterial growth (Monod
1942). Monod’s function of microbial growth is structurally equivalent to the Holling
type II function and it has been used with enormous success during the last 60 years.
However, other functions have been discussed in microbiology and, most interestingly
for this study, Contois published already in 1959 a growth function for microorganisms
that is equivalent to the particular ratio-dependent model used in this study. Many
authors have used Contois’ function without comparison to Monod’s function and they
obtained good …ts to their data (mostly sewage and fermentation processes) (Bala &
Satter 1990; Tijero et al. 1989; Lequerica et al. 1984; Pareilleux & Chaubet 1980;
Ghaly & Echiegu 1993). Table 8.3 lists studies that compared several functions to data
according to a single selection criterion. While Contois’ function often …tted better than
Monod’s function, model selection was also often ambiguous (as in the present work).
Many of these results deal with systems in which the prey or the predator are sets of
many species. While monospeci…c systems seem to be better approximated by Monod’s
model (Grady Jr. et al. 1972), just as we saw in our analysis with Gause’s, Huffaker’s
and Luckinbill’s data, multispecies systems as cited above and in Table 8.3 seem to
Jost, Arditi
Identifying predator-prey models II
135
favour predator-dependent models (Grady Jr. & Williams 1975; Elmaleh & Ben Aim
1976; Daigger & Grady Jr. 1977).
Table 8.3: Collection of studies that compared Contois’ function quantitatively
with other functions. The models tested are noted in the third column with the
best …tting model, if there was one, in capitals. References are: (1) Chiu et al.
(1972), (2) Ashby (1976), (3) Morrison et al. (1987), (4) Dercová et al. (1989),
(5) Wilhelm (1993).
ref. system
tested functions
(1) microbial sewage
Moser, Monod, Contois
(2) protozoa feeding on bacteria
Contois, Ashby (Monod’s
function divided by predator
density)
(3) nutrient limited phytoplankton growth
Monod, Contois, logistic
(4) growth and glucose consumption of yeast Contois, Monod
(5) protozoa feeding on bacteria
Contois, Monod and 9 others
?
?
best …ts were obtained with functions that are sigmoid with respect to substrate
concentration, either Contois or Monod type
To summarise, systems with monospeci…c prey and predator show, in general, little
predator dependence in the functional response except in cases where predators have
a strong potential to interfere with each other (e.g., cannibalism). More complex systems such as plankton in freshwater lakes show a multitude of patterns. We found
no indicators in these systems that tell whether predator dependence in the functional
response should be included or not. We conclude that whenever predictions must be
done by model simulation, at least two different models should be used to distinguish
robust features from model-dependent features. Parameters of such models should be
estimated as well as possible by direct measurements in the …eld, with nonlinear regression being used only for …ne-tuning. The …tting itself should account for stochasticities
in the observed data, but the techniques used in this study (observation-error-…t and
process-error-…t) leave much room for further improvement.
Acknowledgements
We thank J. P. Pelletier and G. Balvay for kindly providing the data from Lake Geneva
and helping with the data description. We also thank G. Harrison for communicating
the data set he obtained from L. Luckinbill and S. Carpenter for kindly providing the
raw plankton data of Paul and Tuesday Lake. This research was supported by the
Swiss National Science Foundation and by the French ‘Programme Environnement, Vie
et Société’ (CNRS).
136
Identifying predator-prey models (PhD Thesis)
C. Jost
Table 8.4: Results for …tting time-series data. The data sets are described by
their length m and their apparent dynamic behaviour (in parenthesis if difficult to
decide): st for stable non trivial equilibrium, ss for strongly stable equilibrium, l
for limit cycle, e for extinction of both populations, pe for extinction of the predator
only and t for transient trajectory. For each type of …tting (process-error-…t or
observation-error-…t) the better …tting model is indicated (p for prey-dependent
and r for ratio-dependent) with criteria X 2 , X L (both with the dynamic behaviour
of the selected model with the …tted parameters), expectation EX 2 and improved
estimate of prediction error (IE ). s is the prediction horizon in process-error-…t.
Signi…cance of model selection is indicated by a † or by a ‡ ; indicates nonsigni…cance due to a low difference between the …ts to each model; the absence of
a superscript indicates non signi…cance due to a low likelihood (see the text for
the employed de…nition of signi…cance).
data characteristics
process-error-…t
observation-error-…t
2
L
m dyn.
s X
X
EX 2 IE
Xω2
XωL
EX 2 IE
name
†
†
†
gause1
18 st
2 r st p st
r
p
p st p st p†
p†
†
†
†
†
†
†
†
gause3
18 l
2 p st p l
p
p
p l
p l
p
p†
gause4
19 l
2 p† st p† st
p†
p†
p† l
p† l
p†
p†
†
†
gauset2a 10 e
1 p l
p l
p
p
p l
p l
p
p†
gauset2c 10 e
1 p l
p† st
r
r
p† l
p† l
p†
p†
gauset2d 10 e
1 p l
p† l
p
p
p l
p† l
p†
p
gauset2e 9 e
1 p l
p l
p
r
p l
p l
r
p
†
†
gauset2f
9 e
1 r e r st, e r
r
r e r e r
r†
†
†
†
luckin1a 35 l
3 p st p st
p
p
p l
p l
p
p†
luckin1b 24 l
3 p st p† st
p†
p†
p† l
p† l
p†
p†
luckin3a 16 l
3 r† st p† st
p†
p†
p† l
p† l
p†
p†
†
†
†
†
†
luckin4a 27 l
3 p st p st
p
p
p l
p l
p
p†
luckin4b 21 l
3 p l
p† l
p†
p†
p† l
p† l
p†
p†
luckin5
62 l
3 p st p† st
p
p
p† l
p† l
p†
p†
†
†
†
†
†
†
veill8
87 l
3 p st p st
p
p
r l
r l
r
r†
veill10
20 l
3 p l
p† st
p†
p†
p† l
p† l
p†
p†
‡
†
‡
‡
†
†
†
‡ynn1b
14 t
4 r t
r t
r
r
r t
r t
r
r†
‡ynn1c
21 st, e
4 r† e r† e
r†
r†
r† e r† e r‡
r‡
‡ynn2b
28 st, e
4 r† e r† e
r†
r†
r† e r† e r†
r†
†
†
†
†
†
†
†
‡ynn2c
28 st, e
4 r e r e
r
r
r e r e r
r†
wilh4.2
16 st, e
4 r st r† e
r
p
p st p† ss p†
p†
†
†
†
wilh4.4
17 st, e
3 r ss r ss
r
r
r st r st r
r
wilh5.27 17 e
4 p st r ss
r
p
p st r† e p
p
†
†
†
wilh5.28 22 st, e
5 r e r e
r
r
p pe p t
p
p†
†
†
†
†
†
wilh5.29 19 st
3 p ss r ss
p
p
r st p ss p
p
wilh5.30 18 st, e
4 p pe r ss
r
p
p† t
p† t
p†
p†
Jost, Arditi
huff11
huff12
huff13
huff14
huff15
huff16
huff17
huff18
huff63-3
huff63-4
paul84
paul85
paul86
paul87
paul88
paul89
paul90
tues84
tues85
tues86
tues87
tues88
tues89
tues90
edPhy86
edPhy87
edPhy88
edPhy89
edPhy90
edPhy91
edPhy92
edPhy93
totPhy86
totPhy87
totPhy88
totPhy89
totPhy90
totPhy91
totPhy92
totPhy93
Identifying predator-prey models II
12
13
10
10
11
11
12
35
58
23
13
17
14
15
15
13
16
11
13
14
15
13
11
19
14
13
13
13
14
12
11
12
14
13
15
14
14
16
16
16
e, l
e
e
e
e, l
e, l
e
l
l
l
(st)
st
(st)
(st)
(l)
l
(l)
st
st
(l)
(l)
(l)
st
(st)
(l)
(st)
(st)
(l)
(st)
(st)
st
st
(st)
(st)
(l)
(l)
st
(l)
(l)
(l)
3
3
2
2
3
3
3
4
4
3
2
2
2
2
2
2
2
2
2
2
2
2
2
2
2
2
2
2
2
2
2
3
3
2
2
2
2
2
2
2
r
p
p
p
r
p
r
r
p
p
r
p
p
p
r
p
p
p
r
r†
p
p†
p
p
r
p
p
p
r
r
p
r
p
p
p
p
p†
p
r†
r†
st
l
l
l
st
l
st, e
st
st
l
ss
st
st
st
ss
ss
st
ss
e
st
st
st
st
ss
ss
st
st
ss
ss
ss
ss
st, e
t
ss
ss
st
st
ss
ss
ss
r
r
p†
p†
p
p
r†
r
p
r†
r
p†
p
p
r
r†
r†
r
r
r
p
p†
p
r
p
p†
p
r
r
p
p†
r†
r†
p†
r
r
r
p
r
r†
st
e
l
l
l
l
st
st
st
st
ss
ss
st
st
ss
st
ss
ss
e
st
pe
st
st
ss
ss
st
l
st
ss
ss
ss
e
ss
ss
ss
ss
ss
ss
ss
ss
r
p
p
p†
r
p
r
r
p
r
r
p
p†
r
r
p
p
p
p
p†
p
p†
r
r
r
p†
p†
p†
r
r
p
r†
p
p
p
p
p†
p
r†
r†
p
p
p
p†
p
p
p
r
p
p
r
p
p†
p
r
p
r
r
p
p†
p
p†
p†
r†
p
p†
p†
r
r
r
r
r†
r
p†
r
r
p†
r
r†
r
r
p
p†
p†
r
p†
p
p
p
p†
p
p
p
p†
r†
=
r
p†
p
p†
p†
p†
p†
p†
p
p†
p†
r
r†
r†
r
r†
r
p
p
p†
p†
p†
p
p†
st
l
l
l
st
l
l
l
st
l
st
ss
st
st
t
st
st
st
l
l
l
st
st
l
st
st
st
st
st
st
st
t
t
st
st
st
st
st
t
l
137
r
p†
p†
p†
p†
p†
p†
p
p†
p†
r†
p†
r†
p†
p†
p†
r
r
p†
r
r
p†
r†
p†
r†
p
p†
r†
r†
r†
r†
r†
r†
p
p†
p†
p†
r
p
p
st
l
l
l
l
l
l
l
st
l
t
st
t
st
st
l
st
t
t
t
t
st
st
t
ss
st
t
t
st
st
st
t
t
t
st
st
st
st
l
l
r
p
p†
p†
r
p†
p
p
p
p†
r†
p
p
p†
r
=
r
p†
p†
p†
p†
p†
p
p†
p
p†
p†
r
r†
r†
r†
r†
r
p†
p
p†
p†
p
r†
p†
r
p
p†
p†
r
p†
p
p
p
p†
p
p
r†
p†
r†
=
r
p†
p†
p†
p
p†
p
p†
p
p
p†
r
r†
r†
p
r†
r
r
p
p†
p†
p†
p†
p
138
Identifying predator-prey models (PhD Thesis)
C. Jost
Appendix
8.A
Residual bootstrapping and IEP E
We use the same notation as introduced in the section ‘Error functions’. The algorithm described below is of the type residual bootstrapping (Efron & Tibshirani 1993).
This approach assumes that the used model is correct and the parameters obtained by
minimising equation Xk2 are used to construct the bootstrap time-series.
Fitting the whole trajectory (with initial condition as a free parameter or nuisance
parameter) is related to simple non-linear curve …tting and the following algorithm is
taken from Efron & Tibshirani (1993). For process-error-…t the same algorithm can be
used with slight modi…cations that are indicated. Let ˆ be the best …tting solution to
the original data, obtained by minimising equation Xk2 , with the residuals
1+s i m
ˆi = log(Yi ) log y(ti ), ˆ ,
and total error eˆ (s = 0 for observation-error-…t). These residuals represent the empirical
distribution function of the residuals. Now the bootstrap estimates are created by the
following algorithm:
1. (for process-error-…t only) Fix initial values yi? = Yi , 1 i s.
2. Calculate the bootstrap data (yi? )1+sim by
log(yi? ) = log(y (ti ), ˆ)) + ?i
1im
where the ?i are a random sample with replacement from the (ˆk )1+sk m (for
predator-prey data, all residuals are thrown in the same pool). For process-error…t s 1 and the bootstrap data are calculated recursively.
3. Estimate ˆ? from the bootstrap data by minimising equation Xk2 , with …nal error
e? .
4. Go to second item and repeat the loop B times (B = 50).
Efron & Tibshirani (1993) suggest that B = 50 200 is in general sufficient for a reliable
bootstrap estimate. Since regression of a differential equation to data is already quite
costly we used the lower value (B = 50), thus allowing to bootstrap on average 1-4
time-series per day. Mean and standard deviation of the error are now calculated in the
usual way from the B estimates e? .
The method of the improved estimate of prediction error (IEP E ) is described in
Efron & Tibshirani (1993) (chapter 17). Let eˆ? be the error with the original data
obtained using the …tted bootstrap parameters ˆ? . The difference eˆ? e? is in general
Jost, Arditi
Identifying predator-prey models II
139
positive and called the ‘optimism’. The improved estimate of prediction error is now
calculated by
"
#
B
1
1 X
IEP E =
eˆ +
(e ˆ?
e?i )
m s
B i=1 ,i
(s = 0 in case of observation-error-…t) and its standard deviation is taken to be the
standard deviation of the optimism’s scaled by the number of predicted data points.
This estimation of prediction error is independent of the estimated residual variance
and the number of parameters …tted, so it is less model dependent than alternatives
such as Cp and BIC statistics (Efron & Tibshirani 1993).
140
Identifying predator-prey models (PhD Thesis)
C. Jost
Appendix A
Collection of predator-prey models
Following the two principles ‘conservation of mass’ and ‘decomposition of any growth
process into birth and death processes’ leads to the general equations for a predator-prey
model:
dx
= F (x) g(x, y )y
dt
dy
= h ((g (x, y)) y (y )y
dt
where F (x) is the prey growth function, g(x, y ) the functional respnse, h(. . . ) the numerical response (predators produced in terms of numbers or biomass as a function
of the consumption) and (y ) is a predator mortality rate. The terms functional and
numerical response were introduced by Solomon (1949).
Predator-prey models that do not follow the ‘conservation of mass’ principle are not
considered in this collection. Popular predator-prey models of this type go back to an
initial idea of Leslie & Gower (1960) who studied the model
x
dx
= r(1
)x axy
K
dt
y
dy
= cy (1 d ).
x
dt
The obvious modi…cation of replacing the Lotka-Volterra predation term ax by a bounded
Holling type II functional response ax/(1 + ahx) has been proposed, analysed and published by both Tanner (1975) and May (1975).
Whenever power laws are used (e.g., xm ), dimensional problems can occur. They
could in principle be remedied by inserting a ‘dummy’ parameter
mu that is of the same
m
dimension as the used state variable (e.g., replace x by uxx
ux ), but this more
complicated notation is never used in practice (see also section 1.2).
A.1
The prey growth function
Often F (x) is decomposed into F (x) = xf (x) where f is called intrinsic growth rate or
per capita growth rate. Following these scheme three types of growth may be identi…ed:
141
142
Identifying predator-prey models (PhD Thesis)
C. Jost
constant f (x) (exponential or Malthusian growth) or linear decrease of f (x) with
increasing x (logistic growth),
maximal intrinsic growth rate at intermediate density (Allee effect),
nonlinear hyperbolic decrease of the intrinsic growth rate (Gompertz Law, a popular model in clinical oncology).
Possible mathematical forms of these growth functions (and others) are listed in table
A.1.
A.2
The functional response
The functional response (prey eaten per predator per unit of time) has found different
functional expressions in the literature. Tables A.2 and A.3 list a little collection of the
different types with their origin.
C. Jost
Appendix: General predator-prey models
143
Table A.1: Simple growth functions found in the literature
F (x)
source
remarks
rx
Malthus (1798)
exponential
or
Malthusian
growth,
considered like a
‘growth-law’
r(1 Kx )x
Verhulst (1838)
logistic growth equation
x r(1 ( K ) )x
Goel et al. (1971),
theta logistic growth
Gilpin
et
al.
(1976), Richards
(1959)
r(( Kx )g 1)x
1g>0
Rosenzweig (1971)
K x
r K+x x
Smith (1963)
pxm qx
von
Bertalanffy
often m = 2/3 is used by an allo(1951)
metric relationship with the organisms weight
2
kx
Szathmáry (1991),
hyperbolic growth for sexual
Eigen & Schuster
reproduction, the population
(1979)
reaches in…nity in …nite time (so1
kt) 1 )
lution x(t) = ( x(0)
√
+k x
Szathmáry (1991),
subexponential growth (found in
Zielinski & Orgel
enzyme-free replication)
(1987)
p
(r a(x b)2 )x
Allee
(1931),
the condition b < r/a must be
Edelstein-Keshet
ful…lled
(1988)
(kx d x2 )x
Volterra (1938)
incorporating the Allee effect in
sexually reproducing organisms
x
x
K
r(1 K )x( 1) K Allee
(1931),
inverse density dependence at
Bazykin (1985)
low densities
Philip (1957), Sza- (k (x) d(x))x kxe x 1 e x is the probability of a
thmáry (1991)
female being fertilized and corresponds to Volterra’s kx x2 ,
(k (x) d(x)) corresponds to Verhulst’s r(1 x/K ).
x ln(x)
Gompertz (1825)
equivalent to e xt x
K
r ln( x )x
Gompertz (1825)
b
(a(1 e x ) c)x
Gutierrez (1992)
E
( YI+
m)x
Schoener (1978)
mechanistic derivation, m is outx
wash, equivalent to Smith (1963)
and to Getz (1984)
144
Identifying predator-prey models (PhD Thesis)
C. Jost
Table A.2: Functional responses that depend on the prey density (x) only
source
remarks
g(x)
Lotka-Volterra
linear functional response without satax
(Nicholson (1933)
uration
in discrete form)
Maynard
Smith
constant response, destabilising
b
(1975) (Thompson
(1924) in discrete
form)
Rosenzweig (1971) cxl (0 < l 1) approx. type II, but without saturation
ax
Holling (1959a)
type II (disc equation), equivalent to
1+ahx
the Monod type function
a
Ivlev (1961)
type II
b(1 e b x )
cx2
)
Watt (1959)
type III, may be generalised by replacb(1 e
ing 2 by n
bxn
Real (1977)
type
III for n > 1. Studied with n = 2
c+xn
by Takahashi (1964)
bcx2
Jost et al. (1973)
multiple saturation model, type III,
b+dx+cx2
also proposed by Hassell et al. (1977)
ba(x+cx2 )
Arditi (1980)
type III, slope at the origin not 0
b+a(x+cx2 )
bx
Andrews (1968)
Monod-Haldane equation, growth inc+x+dx2
hibition at high densities
adx
Tostowaryk (1972)
type
II at low densities, decline at high
d+a(x+cx3 )
densities
ax
Daphnia feeding on (‘inedible’) …lSokol & Howell
1+bx2
(1981)
amenteous algae, simpli…ed MonodHaldane equation
C. Jost
Appendix: General predator-prey models
145
Table A.3: Functional response models that depend on prey (x) and predator (y )
density
source
g(x, y )
remarks
m
Hassell & Varley
pxy
modi…cation of Lotka-Volterra or
(1969)
Nicholson
log y
Gomatam (1974)
px y
no saturation, inspired by Gompertz
(1825)
l
m
cx y
Strebel & Goel
modi…cation of Rosenzweig (1971), no
(1973)
saturation
pbxy m
Hassell & Rogers
combination of Hassell & Varley
b+px
(1972)
(1969) with disc equation, affects saturation level
g1 (x) yys
g1 (x) is any function from Table A.2.
Rogers & Hassell
(1974)
Affects saturation level. ys is the solution of mys2 + (1 m)ys y = 0.
g1 (x) 1+1y
g1 (x) is any function from Table A.2.
Harrison (1995)
px
Beddington (1975)
generalisation of the disc equation
1+ahx+m(y 1)
px
DeAngelis et al.
equivalent to the preceding for y >> 1.
1+cx+my
(1975)
x y m
Arditi & Akçakaya
combines completely Hassell & Varley
1+hx y m
(1990),
(Suther(1969) with Hollings disc equation
land 1983)
x
Arditi & Michaltype II like ratio-dependent, equivay+hx
ski (1995), Contois
lent to preceding model for m = 1.
(1959)
Watt (1959)
b 1
b 1
Arditi et al. (1978)
p xy
Watt (1959)
Aiba et al. (1968)
Fukaya
et
al.
(1996)
e
e
p
xy
b
m
cx2 y
m
if xp < ya
a if xp > ya
Ksx+x e ky
xm
y+hxm
modi…cation of Ivlev (1961)
type III
type I ratio-dependent form with saturation
used in fermentation processes
146
A.3
Identifying predator-prey models (PhD Thesis)
C. Jost
The numerical response
The numerical response is usually considered to be proportional to the functional response, h ( g(x, y )) = eg (x, y ), with ecological efficiency e and any g(x, y ) from Tables
A.2 or A.3. Other forms are possible but rarely used in the literature (e.g., Arditi et al.
(1978), a ratio-dependent model with a ‘hunger phase’ at low ratios x/y ).
A.4
Predator mortality
Formulations of the predator mortality are summarized in table A.4. See Edwards &
Brindley (1996) for a discussion and references to these forms.
Table A.4: Predator mortality functions
(y ) remarks
constant mortality rate, used since the
Lotka-Volterra predator prey model
Steele & Henderson
results in a slanted predator isocline
y
(1981)
(no paradox of enrichment), justi…ed
with increased predation by higher
predators
y
Steele & Henderson
more realistic response to increased
b+y
(1992)
higher predation
DeAngelis
et
al. s + y
(1975),
Bazykins
model
source
Appendix B
Data from Lake Geneva
These are the detailed time-series from Lake Geneva that were used in chapter 8. The
data have been provided by J. P. Pelletier and G. Balvay from the hydrological station
at Thonon, INRA. The temperature and phosphorous data were not directly used in
the …tting but are added here for completeness.
Observed state variables and their units are the following:
State variable
Nano phytoplankton
Total phytoplankton
Herbivorous zooplankton
Temperature at 5m water depth
Dissolved Phosphorous PO4
at 5m water depth
abbreviation
NanP
TotP
HerbZ
T5
unit
mg/m3 fresh weight
mg/m3 fresh weight
mg/m3 fresh weight
C
PO4
g/l
147
148
Identifying predator-prey models (PhD Thesis)
C. Jost
Plankton dynamics year 1986
Date
21-Jan-86
18-Feb-86
3-Mar-86
17-Mar-86
8-Apr-86
21-Apr-86
5-May-86
20-May-86
2-Jun-86
16-Jun-86
7-Jul-86
21-Jul-86
4-Aug-86
18-Aug-86
8-Sep-86
23-Sep-86
6-Oct-86
6-Nov-86
17-Nov-86
8-Dec-86
NanP
131.6
89.7
28.7
361.8
438.1
521.1
1053.0
661.4
101.0
50.1
805.9
TotP
380.6
415.4
180.7
444.5
494.1
556.6
1063.8
675.5
139.1
62.8
825.1
1522.3
762.8
1712.0
2015.6
621.6
239.0
317.3
225.1
3046.4
1887.9
2225.0
2129.6
685.7
346.7
497.4
329.3
HerbZ
334
146
176
316
370
458
1122
2568
4906
798
6518
980
1746
1184
1076
2740
2044
614
190
476
T5
5.9
5.2
5
5.3
5.6
5.8
7.9
9.6
13.8
12.8
16.9
16
20
20.1
17.9
17.3
14.5
11.4
10.7
9
PO4
50
61
64
59
58
59
46
3
13
32
10
8
2
3
4
2
4
15
10
22
7000.0
6000.0
5000.0
4000.0
Nano Phytopl.
Total Phytopl.
Herb. Zoopl.
3000.0
2000.0
1000.0
Figure B.1: Plankton dynamics year 1986
8-Dec
17-Nov
6-Nov
6-Oct
23-Sep
8-Sep
18-Aug
4-Aug
21-Jul
7-Jul
16-Jun
2-Jun
20-May
5-May
21-Apr
8-Apr
17-Mar
3-Mar
18-Feb
21-Jan
0.0
C. Jost
Appendix: Data Lake Geneva
149
Plankton dynamics year 1987
Date
26-Jan-87
9-Feb-87
2-Mar-87
17-Mar-87
6-Apr-87
21-Apr-87
11-May-87
18-May-87
2-Jun-87
22-Jun-87
6-Jul-87
22-Jul-87
3-Aug-87
18-Aug-87
7-Sep-87
22-Sep-87
5-Oct-87
20-Oct-87
2-Nov-87
16-Nov-87
7-Dec-87
NanP
184.7
157.1
101.5
123.0
581.6
2264.6
2359.6
3547.1
948.8
129.1
861.1
337.5
620.2
503.0
851.4
846.1
542.7
633.2
588.0
248.8
191.1
TotP
257.6
217.5
132.1
299.9
934.6
3289.8
3281.5
4185.1
991.7
236.2
2288.8
2014.0
7568.7
2125.5
1626.4
1191.5
846.2
849.3
692.3
310.2
256.6
HerbZ
170
96
388
244
196
324
264
538
340
2548
406
1276
666
498
2242
2172
182
992
288
336
208
T5
5.4
5.1
5.4
5.2
5.4
8.2
8.8
9.8
11.1
14
13.1
16.7
19.1
16.3
19.9
20.5
15
13.2
12.9
11.2
9
PO4
50
56
52
54
52
35
34
12
4
20
6
9
8
3
3
5
2
2
1
8
18
8000.0
7000.0
6000.0
5000.0
Nano Phytopl.
Total Phytopl.
4000.0
Herb. Zoopl.
3000.0
2000.0
1000.0
Figure B.2: Plankton dynamics year 1987
7-Dec
16-Nov
2-Nov
20-Oct
5-Oct
7-Sep
22-Sep
18-Aug
3-Aug
22-Jul
6-Jul
2-Jun
22-Jun
18-May
11-May
21-Apr
6-Apr
2-Mar
17-Mar
9-Feb
26-Jan
0.0
150
Identifying predator-prey models (PhD Thesis)
C. Jost
Plankton dynamics year 1988
Date
11-Jan-88
15-Feb-88
14-Mar-88
29-Mar-88
11-Apr-88
25-Apr-88
9-May-88
24-May-88
8-Jun-88
20-Jun-88
5-Jul-88
19-Jul-88
3-Aug-88
16-Aug-88
5-Sep-88
20-Sep-88
4-Oct-88
17-Oct-88
9-Nov-88
28-Nov-88
12-Dec-88
NanP
142.4
322.5
171.9
50.4
676.6
1350.5
1636.7
109.1
56.3
327.7
241.0
311.6
849.5
395.1
224.1
185.4
227.1
229.5
128.2
208.4
198.7
TotP
300.2
495.2
291.0
149.9
3087.9
2993.4
1757.6
160.6
83.2
814.1
562.0
3529.5
1909.4
2187.7
3716.2
2997.8
2608.5
1696.7
433.4
619.2
332.8
HerbZ
92
214
152
110
120
310
476
1106
1528
774
1146
534
170
262
274
584
74
210
154
244
284
T5
7.2
6.2
5.6
6.1
7.3
9.6
12.8
13.1
14.3
17.6
18.5
20.4
21.6
16.2
18.8
16.2
15
14.6
11.3
8.8
7.8
PO4
34
42
44
46
20
18
5
3
4
2
3
3
3
1
2
4
2
4
11
22
26
4000.0
3500.0
3000.0
2500.0
Nano Phytopl.
Total Phytopl.
2000.0
Herb. Zoopl.
1500.0
1000.0
500.0
Figure B.3: Plankton dynamics year 1988
12-Dec
28-Nov
9-Nov
17-Oct
4-Oct
5-Sep
20-Sep
16-Aug
3-Aug
19-Jul
5-Jul
8-Jun
20-Jun
24-May
9-May
25-Apr
11-Apr
29-Mar
14-Mar
15-Feb
11-Jan
0.0
C. Jost
Appendix: Data Lake Geneva
151
Plankton dynamics year 1989
Date
16-Jan-89
13-Feb-89
6-Mar-89
20-Mar-89
10-Apr-89
24-Apr-89
9-May-89
22-May-89
5-Jun-89
19-Jun-89
5-Jul-89
24-Jul-89
7-Aug-89
21-Aug-89
11-Sep-89
19-Sep-89
2-Oct-89
16-Oct-89
6-Nov-89
20-Nov-89
11-Dec-89
NanP
34.1
26.4
46.6
99.1
458.0
2376.8
1270.0
1206.3
134.8
749.8
522.0
353.5
271.0
389.4
363.0
1460.9
421.8
436.8
139.5
281.4
58.9
TotP
150.7
94.9
297.3
630.5
843.9
3533.8
1443.1
1225.2
180.1
869.0
1042.5
1624.5
2329.1
1486.1
902.6
1820.7
698.6
690.4
169.2
388.8
138.5
HerbZ
76
104
266
290
160
610
434
1066
3032
1370
1934
540
322
592
380
1688
2086
818
982
132
30
T5
6.7
6.1
6.4
6.8
7.7
9
10.5
11.7
16.6
16.4
17
17.7
21.1
22.5
18
18.5
15.5
13.2
12.8
10.5
6.9
PO4
32
42
36
34
24
3
3
6
0
0
2
1
2
2
3
5
3
3
3
7
35
4000.0
3500.0
3000.0
2500.0
Nano Phytopl.
2000.0
Total Phytopl.
Herb. Zoopl.
1500.0
1000.0
500.0
Figure B.4: Plankton dynamics year 1989
11-Dec
20-Nov
6-Nov
16-Oct
2-Oct
19-Sep
11-Sep
21-Aug
7-Aug
24-Jul
5-Jul
5-Jun
19-Jun
22-May
9-May
24-Apr
10-Apr
6-Mar
20-Mar
13-Feb
16-Jan
0.0
152
Identifying predator-prey models (PhD Thesis)
C. Jost
Plankton dynamics year 1990
Date
15-Jan-90
20-Feb-90
5-Mar-90
2-Apr-90
12-Apr-90
25-Apr-90
9-May-90
21-May-90
6-Jun-90
18-Jun-90
9-Jul-90
25-Jul-90
6-Aug-90
20-Aug-90
3-Sep-90
18-Sep-90
1-Oct-90
15-Oct-90
13-Nov-90
26-Nov-90
17-Dec-90
NanP
216.5
212.9
177.5
689.8
1179.1
3541.4
1847.2
243.7
399.6
655.3
474.2
481.1
478.3
226.9
389.1
296.3
110.8
119.7
179.2
122.9
174.2
TotP
282.2
388.9
416.2
1060.6
1483.5
3702.2
1894.2
271.6
771.1
1135.0
1381.3
1689.0
1455.6
1050.9
1589.7
1437.5
329.8
220.2
229.0
196.3
281.2
HerbZ
68
162
80
104
258
498
958
1994
1690
612
562
518
396
414
432
616
474
1044
288
314
174
T5
7
6.9
6.6
7.2
7.3
8.8
14.6
16
16.6
17.3
19.4
16
22.2
21.9
19.8
18.7
17.1
15.4
11.4
10.3
7.5
PO4
25
29
33
27
21
16
6
3
5
3
2
0
3
2
4
3
3
3
6
7
21
4000.0
3500.0
3000.0
2500.0
Nano Phytopl.
Total Phytopl.
2000.0
Herb. Zoopl.
1500.0
1000.0
500.0
Figure B.5: Plankton dynamics year 1990
17-Dec
26-Nov
13-Nov
15-Oct
1-Oct
3-Sep
18-Sep
20-Aug
6-Aug
25-Jul
9-Jul
6-Jun
18-Jun
21-May
9-May
25-Apr
12-Apr
2-Apr
5-Mar
20-Feb
15-Jan
0.0
C. Jost
Appendix: Data Lake Geneva
153
Plankton dynamics year 1991
Date
23-Jan-91
18-Feb-91
4-Mar-91
18-Mar-91
8-Apr-91
22-Apr-91
14-May-91
27-May-91
10-Jun-91
24-Jun-91
8-Jul-91
22-Jul-91
5-Aug-91
26-Aug-91
9-Sep-91
25-Sep-91
8-Oct-91
29-Oct-91
18-Nov-91
16-Dec-91
NanP
130.5
78.4
1097.2
312.0
1874.0
1246.4
2631.5
694.6
293.0
537.8
540.0
636.7
788.2
225.7
160.1
204.6
368.8
342.5
144.7
282.5
TotP
224.0
142.7
1220.9
431.4
1929.0
1270.8
2643.8
705.5
336.2
630.7
935.6
2377.0
1222.7
1101.2
508.1
723.6
662.1
1389.0
398.7
397.1
HerbZ
150
110
128
272
780
994
618
2280
1044
376
136
132
256
292
510
228
370
410
62
T5
6.5
5.9
6
7.3
7.2
7.1
8.9
10.9
15.7
15.9
20
20.3
21.3
22.9
19.5
18.7
15.7
11.3
10.2
7
PO4
28
34
29
8
8
16
13
6
16
5
2
3
4
6
5
4
3
3
7
22
3000.0
2500.0
2000.0
Nano Phytopl.
Total Phytopl.
1500.0
Herb. Zoopl.
1000.0
500.0
Figure B.6: Plankton dynamics year 1991
16-Dec
18-Nov
29-Oct
8-Oct
25-Sep
9-Sep
26-Aug
5-Aug
22-Jul
8-Jul
24-Jun
10-Jun
27-May
14-May
22-Apr
8-Apr
18-Mar
4-Mar
18-Feb
23-Jan
0.0
154
Identifying predator-prey models (PhD Thesis)
C. Jost
Plankton dynamics year 1992
Date
15-Jan-92
24-Feb-92
2-Mar-92
17-Mar-92
7-Apr-92
21-Apr-92
4-May-92
20-May-92
1-Jun-92
15-Jun-92
6-Jul-92
20-Jul-92
3-Aug-92
17-Aug-92
7-Sep-92
22-Sep-92
14-Oct-92
27-Oct-92
9-Nov-92
23-Nov-92
14-Dec-92
NanP
151
124
253
282
705
601
1890
2117
193
932
562
216
348
1028
72
328
204
60
81
132
68
TotP
209
177
494
354
754
683
1929
2181
292
978
1190
634
2425
3334
1265
2887
1445
1946
1091
988
214
HerbZ
158
272
218
438
270
512
856
620
1822
1030
1058
410
322
334
450
442
586
236
228
244
328
T5
6.39
5.6
6.17
6.17
6.45
7.5
9.83
9.32
15.45
15.5
17.5
19.1
16.52
21.58
18.37
19.12
14.73
11.7
10.97
9.79
8.42
PO4
25
24
23
23
18
27
4
1
5
4
3
2
3
3
3
3
2
1
3
10
13
3500
3000
2500
2000
Nano Phytopl.
Total Phytopl.
Herb. Zoopl.
1500
1000
500
Figure B.7: Plankton dynamics year 1992
14-Dec
23-Nov
9-Nov
27-Oct
14-Oct
22-Sep
7-Sep
17-Aug
3-Aug
6-Jul
20-Jul
15-Jun
1-Jun
20-May
4-May
21-Apr
7-Apr
17-Mar
2-Mar
24-Feb
15-Jan
0
C. Jost
Appendix: Data Lake Geneva
155
Plankton dynamics year 1993
Date
14-Jan-93
2-Mar-93
15-Mar-93
5-Apr-93
19-Apr-93
3-May-93
17-May-93
7-Jun-93
21-Jun-93
5-Jul-93
26-Jul-93
2-Aug-93
23-Aug-93
6-Sep-93
21-Sep-93
4-Oct-93
3-Nov-93
8-Nov-93
24-Nov-93
NanP
232.9
194.9
387.9
571.4
2985.6
2780.0
2063.0
858.5
341.7
493.5
208.1
85.2
63.7
122.4
110.9
149.8
405.9
848.6
159.1
TotP
445.2
708.4
591.7
777.2
3315.8
2983.0
2155.6
935.1
590.0
909.8
3685.1
3289.3
1404.0
2485.0
1243.0
2173.5
1513.6
1679.0
382.0
HerbZ
364
170
232
210
164
1366
2000
1278
572
350
374
368
442
242
592
528
128
114
482
T5
PO4
4000.0
3500.0
3000.0
2500.0
Nano Phytopl.
Total Phytopl.
2000.0
Herb. Zoopl.
1500.0
1000.0
500.0
Figure B.8: Plankton dynamics year 1993
24-Nov
8-Nov
3-Nov
4-Oct
21-Sep
6-Sep
23-Aug
2-Aug
26-Jul
5-Jul
21-Jun
7-Jun
17-May
3-May
19-Apr
5-Apr
15-Mar
2-Mar
14-Jan
0.0
156
Identifying predator-prey models (PhD Thesis)
C. Jost
Appendix C
Distinguishability and identi…ability
of the studied models
C.1
Distinguishability
Let us …rst de…ne what we mean by distinguishability (sensu Walter & Pronzato (1997)):
ˆ ()
ˆ , with their vector of parameters and ˆ
if there are two models, M () and M
ˆ (.) if for almost any
respectively, we say M (.) is structurally distinguishable from M
ˆ
ˆ
realization of parameters for M (.) there is no realization for M (.) such that M () =
ˆ ()
ˆ .
M
Consider we have a set of non-zero parameters for the prey-dependent model. Equating the prey equations of both models we get
r(1
N
)N
K
aN P
= (1
1 + ahN
N
)N
NP
P + N
∀N =
6 0, P =
6 0.
Multiplying both sides by (1 + ahN )(P + N ) we get a polynomial equation in N
and P . The coefficients of a polynom are unique, therefore the coefficients belonging
to any N i P j must be the same on the left and on the right side. Since the term aP 2
has no counterpart on the right side this is only possible for a = 0 in violation of the
assumption of non-zero parameters. We may therefore conclude that the two models
are distinguishable.
C.2
Identi…ability
Of the various methods described in (Walter 1987) we will apply here the Taylor series
expansion for the case of exact observations without error. The latter hypotheses enables
us in principle to calculate successive derivatives of the population abundances, thus
 (0) may be considered as known constant values.
expressions like N˙ (0) or N
In the Taylor series expansion approach the outputs (abundances in time, trajectories) are developed in a Taylor series about t = 0+ of which the successive terms can be
157
158
Identifying predator-prey models (PhD Thesis)
C. Jost
calculated and can be expressed as functions of the unknowns. Thus we can obtain a
sufficient number of equations that can be solved for the unknown parameters.
C.2.1
Predator-prey model with prey-dependent functional response
Consider the model
N
)N
K
aN
P˙ = e
P
1 + ahN
N˙ = r(1
aN
P,
1 + ahN
P,
N (0) = N0
P (0) = P0 .
This model has 6 parameters that need to be estimated, so we will need at least 6
equations.
N˙
N
aP
(C.1)
= r(1
)
=: N
K
1 + ahN
P˙
aP
(C.2)
=: =e
P
1 + ahN

NN
N˙ 2 (N˙ =N,P˙ =P ) rN
a2 hNP
aP
(C.3)
∂t =
=
+
=: 2
2
N
K
(1 + ahN )
1 + ahN
P P P˙ 2 (N˙ =N,P˙ =P )
eaN
(C.4)
∂t =
=
=: 2
P
(1 + ahN )2
a2 hNP (2 + ( 2 + )) aP ( 2 + )
( 2 + )rN
2a3 h2 2 N 2 P
(C.5)
∂tt =
+
K
(1 + ahN )3
(1 + ahN )2
1 + ahN
2
( + )N
2ah 2 N 2
(C.6)
∂tt = ae
(1 + ahN )2
(1 + ahN )3
 = N ( + 2 ) and P = P ( +
The last two equations were simpli…ed using the relations N
2 ) (see equations (C.3) and (C.4)). We can further simplify equations (C.3,C.5,C.6)
1
by using relation (C.4): (1+ahN
. Since only appears in equation (C.2) we can
= eaN
)2
solve the remaining …ve equations for the remaining parameters and calculate at the
end.
N˙
N
aP
(C.7)
= r(1
)
=: N
K
1 + ahN
rN
ahP
aP
(C.8)
∂t =
+
=: K
e
1 + ahN
eaN
(C.9)
=: ∂t =
(1 + ahN )2
( 2 + )rN
2a2 h2 NP
∂tt =
(1 + ahN )e
K
aP ( 2 + )
ahP (2 + ( 2 + ))
(C.10)
=: +
1 + ahN
e
( 2 + )
2ahN
=: (C.11)
∂tt =
1 + ahN
C. Jost
Appendix: Distinguishability, Identi…ability
159
Evaluating all …ve new equations at t = 0 we obtain …ve equations with …ve unknown
parameters and can now test if there exists a unique positive solution for them. This
involves algebraic manipulations which can be done in any software such as Mathematica
or Maple, so here only the main steps will be described, not the intermediate results.
The parameters r and K only appear in equations (C.7,C.8) and (C.10). Solve the
…rst two equations for r and K and insert the (unique) solution into equation (C.10).
Now solve equation (C.9) for h (there will be only one solution that is positive) and
insert it into (C.11) and into the updated version of (C.10), which build now a system of
two equations in the unknowns e and h. Solve now the updated version of (C.11) for e,
plug it into (C.10), which will thus become a polynomial of …rst degree in a. Therefore,
there exists a unique solution that can be positive in all parameters. Finally, plug the
solutions for a and h back into (C.2) to obtain .
C.2.2
Predator-prey model with ratio-dependent functional
response
Consider the model
N
N
)N
P,
N (0) = N0
K
P + hN
N
P˙ = e
P P,
P (0) = P0 .
P + hN
N˙ = r(1
We will proceed in the same manner as in the last section. The six basic equations
become:
N˙
N
P˙
P
N
P
)
=: K
P + hN
P
e
=: P + hN

NN
N˙ 2 (N˙ =N,P˙ =P ) rN
2 NP h( )
=
+
=: (P + hN )2
N2
K
P P P˙ 2 (N˙ =N,P˙ =P )
eN P
=
=: 2
(P + hN )2
P
( 2 + )rN
2P (hN + P )2 P ( 2 + )
(P + hN )3
K
P + hN
2
2P (hN + P ) + P (h( + )N + ( 2 + )P )
+
(P + hN )2
eNP ( ) eN P ( )2 (P hN )
.
(P + hN )2
(P + hN )3
= r(1
(C.12)
=
(C.13)
∂t =
∂t =
∂tt =
∂tt =
(C.14)
(C.15)
(C.16)
(C.17)
 = N ( + 2 ) and P =
The last two equations were simpli…ed using the relations N
2
P ( + ) (see equations (C.14) and (C.15)). We can further simplify equations (C.14,
C.16, C.17) by using relation (C.15): 1/(P + hN )2 = /(eNP ( )). Since only
Identifying predator-prey models (PhD Thesis)
160
C. Jost
appears in equation (C.13) we can solve the remaining …ve equations for the remaining
parameters and calculate at the end.
N˙
N
= r(1
N
)
K
P
=: P + hN
rN
h
+
=: K
e
eNP ( )
=: ∂t =
(P + hN )2
( 2 + )rN
2(hN + P )2
P ( 2 + )
∂tt =
(P + hN )eN ( )
P + hN
K
2
2
2 (hN + P ) + h( + )N + ( + )P
+
=: eN ( )
( )
( )(P hN )
∂tt =
+
=: ( )
P + hN
∂t =
(C.18)
(C.19)
(C.20)
(C.21)
(C.22)
Evaluating all …ve new equations at t = 0 we obtain …ve equations with …ve unknown
parameters and can now test if there exists a unique positive solution for them.
Solve equations (C.18) and (C.19) for parameters r and K and insert the solutions
into equation (C.21). Solve now equation (C.20) for h (there is at most one positive
solution) and insert it into equation (C.22) and into the updated version of equation
(C.21). Now solve this new version of equation (C.22) for e and insert it into (the new)
equation (C.21). This has now …nally become a polynomial of …rst degree in and thus
there exists at most one positive solution of the equations (C.12) to (C.17).
Appendix D
Transient behavior of general 3-level
trophic chains
In a general food chain of length n the rate of change of each state variable depends
both on lower and higher state variables:
x˙ 1 = f1 (x1 , x2 )
..
.
x˙ i = fi (xi 1 , xi , xi+1 )
..
.
x˙ n = fn (xn 1 , xn ).
2i<n
Bernard & Gouzé (1995) presents a method for determining the possible succession
of maxima and minima of the state variables in a food chain where the rate of change
of each state variable only depends on itself and the next higher state variable. With
slight modi…cations (personal communication with the authors of Bernard & Gouzé
(1995)), this method may be applied to food chains, de…ned as above, and thus one
can determine the possible succession of maxima and minima. This method is applied
to study general prey-dependent and ratio-dependent 3-level food chains with regard
to their ability to describe the succession of events in the PEG-model (Sommer et al.
1986).
D.1
The method
The method is based on studying the sign-transitions (i.e. changes of the sign of the
derivative) of state variables in the velocity space instead of the phase space. Let
zi (t) := x˙ i (t) = fi (xi
1 , xi , xi+1 ),
(in a physical analogy, if x(t) is abundance at time t, z(t) is the velocity of the abundance
at time t), then we can calculate the acceleration
z˙i (t) = ti zi
1
+ di zi + si zi+1
161
Identifying predator-prey models (PhD Thesis)
162
C. Jost
with ti = ∂xi 1 fi , di = ∂xi fi and si = ∂xi+1 fi . If the state variable xi is at an extremum,
then zi = 0 and with knowledge of the signs of the velocities before the extremum (i.e.
if zi 1 and zi+1 were positive or negative) and knowledge of the signs of ti and si we
may determine if xi changed from a negative slope to a positive slope (xi at minimum)
or vice versa. Therefore we can create all possible successions of extrema of our system
and compare them with real data in order to reject the model if it cannot predict the
observed pattern.
We won’t dwell here on the mathematical details, but simply mention a necessary
condition in addition to the ones given in Bernard & Gouzé (1995):
The set of trajectories such that two state variables admit at the same time an
extremum is of zero measure.
For the trophic chains that we will consider here this and the other conditions Bernard
& Gouzé (1995) are always ful…lled.
D.2
The 3-level trophic chains and the PEG-model
Consider the model
x1
)x1 g1 (x1 , x2 )x2
K
= f2 (x1 , x2 , x3 ) = e1 g1 (x1 , x2 )x2 g2 (x2 , x3 )x3
= f3 (x2 , x3 ) = e2 g2 (x2 , x3 )x3 x3
x˙ 1 = f1 (x1 , x2 ) = r(1
x˙ 2
x˙ 3
where g1 and g2 are either prey-dependent (left) or ratio-dependent (right) functional
responses,
g(x1 )
g(x2 )
x1
)
x2
x2
g2 (x2 , x3 ) → g( ).
x3
g1 (x1 , x2 ) → g(
In general, g1 and g2 are positive, monotonically increasing, concave functions of their
i
respective arguments, gi (y ) > 0, g˙i (y ) > 0 and gi (y ) < 0 for y = xi or y = xix+1
(for a
justi…cation of this assumption see Arditi & Ginzburg (1989)) and gi (0) = 0.
In the context of the PEG-model x1 is the phytoplankton, x2 the herbivorous zooplankton and x3 carnivorous zooplankton or planktivorous …sh. Succession of maxima
and minima are given for the phytoplankton and herbivorous zooplankton: both populations start with increasing abundances (z1 (0) > 0, z2 (0) > 0), phytoplankton shows
the …rst maximum, followed by a zooplankton maximum, the phytoplankton shows the
spring depression (clear water phase), followed by a minimum in the zooplankton. Then
phytoplankton reaches the next maximum again before the zooplankton. If the signs of
the zi are written in a vertical box, this succession may be written as
C. Jost
Appendix: Transient behaviour
163
’
+
+
-
- +
-
-
+
+
- +
$
sign(z1 )
sign(z2 )
sign(z3 )
-
- + :=
&
D.2.1
%
Analysis of the prey-dependent model
We have the following signs in the jacobian:
t1
t2
t3
s1
s2
s3
=
=
=
=
=
=
∂x3 f1
∂x1 f2
∂x2 f3
∂x2 f1
∂x3 f2
∂x1 f3
=0
= e1 x2 ∂x1 g1 (x1 ) > 0
= e2 x3 ∂x1 g1 (x1 ) > 0
= g1 (x1 ) < 0
= g2 (x2 ) < 0
=0
Therefore, the possible transitions of the zi , depending on the signs of zi 1 and zi+1 , are
shown in …gure D.1. If z1 z3 > 0 then z2 can evolve in both ways (marked as p for
‘possible’ in the transition graph in …gure D.2).
for s1 < 0, t1 = 0
for s2 < 0, t2 > 0
+
-
for s3 = 0, t3 > 0
+
-
m1
+
- or
+
+
or
+
+
or
+
m2
+
- +
-
m3
- +
+
M1
-
- +
M2
-
M3
-
+
-
Figure D.1: The possible transitions between extrema for the prey-dependent and
the ratio-dependent model.
We may now study all possible transitions of the whole system (see …gure D.2). We
can conclude that our model can predict the succession of events in the PEG-model.
164
Identifying predator-prey models (PhD Thesis)
p
:
+
+
+
M1
- +
+
:
:
:
m1
+ M3
+ @
@
@R
M2
- M3
+ @
m1
+
-
@
@R -
-
+
+
-
- +
-
m3
- +
+
XX
XX
M1
-
- +
:
:
-
+ M3
+ @
- M3
+ @
p
+
- +
-
p
M2
XXX
XXX
m2
:
:
m1
M1
C. Jost
@
@R
m1
@
@R -
XX
XX
Xz -
:
+
-
m2
+
- +
:
:
M1
-
- +
:
-
p
Figure D.2: The possible transitions for the prey- and ratio-dependent model.
The straight arrows refer to transitions consistent with the PEG-model, while the
dashed arrows refer to other transitions. A p indicates alternative possibilities.
D.2.2
Analysis of the ratio-dependent model
We have the following signs in the jacobian:
t1 = ∂x3 f1 = 0
x1
)>0
x2
x2
= e2 g˙2 ( ) > 0
x3
x1 x1
x1
= ( g˙ 1 ( ) + g1 ( )) < 0
x2 x2
x2
x2 x2
x2
= ( g˙ 2 ( ) + g2 ( )) < 0
x3 x3
x3
= 0.
t2 = ∂x1 f2 = e1 g˙1 (
t3 = ∂x2 f3
s1 = ∂x2 f1
s2 = ∂x3 f2
s3 = ∂x1 f3
(D.1)
(D.2)
Inequalities (D.1) and (D.2) hold because of the Mean Value Theorem (see below).
Since all the signs are the same as in the prey-dependent model the transitions will be
the same as in …gure D.1 and in …gure D.2 and so the ratio-dependent model may also
C. Jost
Appendix: Transient behaviour
165
predict the events of the PEG-model.
Proof of inequalities (D.1) and (D.2) Let f be a strongly concave function (f < 0
on the interval I := [0, ∞) with f (0) = 0. The Mean Value Theorem states that for any
x, y ∈ I, x < y there exists a ∈ (x, y ) such that f (y ) f (x) = (y x)f˙( ). Since f is
concave, f˙( ) > f˙(y ) and so we get the inequality
f (y) > f (x) + (y
With f = gi and y =
xi
xi+1
x=0
x)f˙(y ) = yf˙(y ).
this proves inequalities (D.1) and (D.2).
q.e.d.
166
Identifying predator-prey models (PhD Thesis)
C. Jost
Bibliography
Abrams, P. A. 1993. Effect of increased productivity on the abundances of trophic
levels. American Naturalist 141:352–371.
Abrams, P. A. 1994. The fallacies of “ratio-dependent” predation. Ecology 75:1842–
1850.
Abrams, P. A. 1997. Anomalous predictions of ratio-dependent models of predation.
Oikos 80:163–171.
Abrams, P. A., & J. D. Roth. 1994. The effects of enrichment of three-species food
chains with nonlinear functional responses. Ecology 75:1118–1130.
Agrawal, A. A. 1998. Algal defense, grazers, and their interactions in aquatic trophic
cascades. Acta Œcologica 19:331–337.
Aiba, S., M. Shoda, & M. Nagatani. 1968. Kinetics of product inhibition in alcohol
fermentation. Biotechnology and Bioengineering 10:845–864.
Akaike, H. 1973. Information theory and an extension of the maximum likelihood
principle. In B. N. Petran, & F. Csaki, eds., International Symposium on Information Theory , pages 267–281. Akademia Kiado, Budapest, Hungary, 2nd edn.
Akçakaya, H. R., R. Arditi, & L. R. Ginzburg. 1995. Ratio-dependent predation: an
abstraction that works. Ecology 76:995–1004.
Allee, W. C. 1931. Animal Aggregations. A Study in General Sociology. University
of Chicago Press.
Amann, H. 1990. Ordinary Differential Equations. de Gruyter, Berlin.
Andrews, J. F. 1968. A mathematical model for the continuous culture of microorganisms utilizing inhibitory substrates. Biotechnology and Bioengineering 10:707–723.
Arditi, R. 1980. Identi…cation de la réponse fonctionnelle de prédateurs et de parasitoïdes. In Actes du 8ème colloque Informatique et Biosphère , pages 333–349,
Paris.
Arditi, R., & H. R. Akçakaya. 1990. Underestimation of mutual interference of predators. Oecologia 83:358–361.
Arditi, R., & A. A. Berryman. 1991. The biological control paradox. Trends in Ecology
and Evolution 6:32.
Arditi, R., & L. R. Ginzburg. 1989. Coupling in predator-prey dynamics: ratiodependence. Journal of Theoretical Biology 139:311–326.
167
168
Identifying predator-prey models (PhD Thesis)
C. Jost
Arditi, R., & J. Michalski. 1995. Nonlinear food web models and their responses to
increased basal productivity. In G. A. Polis, & K. O. Winemiller, eds., Food Webs:
Integrations of Patterns and Dynamics , pages 122–133. Chapman & Hall.
Arditi, R., & H. Saïah. 1992. Empirical evidence of the role of heterogeneity in ratiodependent consumption. Ecology 73:1544–1551.
Arditi, R., J. M. Abillon, & J. Vieira da Silva. 1978. A predator-prey model with
satiation and intraspeci…c competition. Ecological Modelling 5:173–191.
Arditi, R., L. R. Ginzburg, & H. R. Akçakaya. 1991a. Variation in plankton densities
among lakes: a case for ratio-dependent predation models. American Naturalist
138:1287–1296.
Arditi, R., N. Perrin, & H. Saïah. 1991b. Functional responses and heterogeneities:
an experimental test with cladocerans. Oikos 60:69–75.
Ashby, R. E. 1976. Long term variations in a protozoan chemostat culture. Journal
of Experimental Marine Biology and Ecology 24:227–235.
Åström, M. 1997. The paradox of biological control revisited: per capita nonlinearities. Oikos 78:618–621.
Azam, F., T. Fenchel, J. G. Field, L. A. Meyer-Reil, & F. Thingstad. 1983. The ecology of water-column microbes in the sea. Marine ecology progress series 10:257–
263.
Bala, B. K., & M. A. Satter. 1990. Kinetic and economic considerations of biogas
production systems. Biological Wastes 34:21–38.
Barford, J. P., & R. J. Hall. 1978. An evaluation of the approaches to the mathematical modelling of microbial growth. Process Biochemistry 11:22–29.
Bastin, G., & D. Dochain. 1990. On-line Estimation and Adaptive Control of Bioreactors. Elsevier.
Bazin, M. J. 1981. Theory of continuous culture. In P. H. Calcott, ed., Continuous
Cultures of Cells , vol. 1, pages 27–62. CRC Press, Boca Raton, Florida.
Bazin, M. J., & P. T. Saunders. 1978. Determination of critical variables in a microbial
predator-prey system by catastrophe theory. Nature 275:52–54.
Bazykin, A. D. 1985. Mathematical Models in Biophysics. Nauka, Moscow.
Beddington, J. R. 1975. Mutual interference between parasites or predators and its
effect on searching efficiency. Journal of Animal Ecology 44:331–340.
Beddington, J. R., C. A. Free, & J. H. Lawton. 1978. Characteristics of successful
natural enemies in models of biological control of insect pests. Nature 273:513–519.
Begon, M., M. Mortimer, & D. J. Thompson. 1996a. Population ecology: a uni…ed
study of animals and plants. Blackwell Science, 3rd edn.
Begon, M., S. M. Sait, & D. J. Thompson. 1996b. Predator-prey cycles with period
shifts between two- and three-species systems. Nature 381:311–315.
Benitez, J., J. Beltran-Heredia, J. Torregrosa, J. L. Acero, & V. Cercas. 1997. Aerobic
degradation of olive mill wastewaters. Applied Microbiology and Biotechnology
47:185–188.
C. Jost
Identifying predator-prey models (PhD Thesis)
169
Beretta, E., & Y. Kuang. 1998. Global analyses in some delayed ratio-dependent
predator-prey sytems. Nonlinear Analysis, Theory, Methods & Applications
32:381–408.
Berman, T. 1990. Microbial food webs and nutrient cycling in lakes: changing perspectives. In M. M. Tilzer, & C. Serruya, eds., Large Lakes
Ecological Structure
and Function . Springer Verlag.
Bernard, O., & J.-L. Gouzé. 1995. Transient behavior of biological loop models, with
application to the Droop model. Mathematical Biosciences 127:19–43.
Bernardi, R. d., & G. Giussani. 1990. Are blue-green algae a suitable food for zooplankton? An overview. Hydrobiologia 200/201:29–41.
Bernstein, C. 1985. A simulation model for an acarine predator-prey system (Phtyoseiulus persimilis - Tetranychus urticae). Journal of Animal Ecology 54:375–389.
Berryman, A. A. 1996. What causes population cycles of forest lepidoptera. Trends
in Ecology and Evolution 11:28–32.
Berryman, A. A., J. Michalski, A. P. Gutierrez, & R. Arditi. 1995. Logistic theory of
food web dynamics. Ecology 76:336–343.
Bilardello, P., X. Joulia, J. M. Le Lann, H. Delmas, & B. Koehret. 1993. A general
strategy for parameter estimation in differential-algebraic systems. Computers and
Chemical Engineering 17:517–525.
Blackman, F. 1905. Optima and limiting factors. Annals of Botany 19:281–295.
Boekel, W. H. M. v., F. C. Hansen, R. Riegman, & R. P. M. Bak. 1992. Lysis-induced
decline of a Phaeocystis spring bloom and coupling with the microbial foodweb.
Marine Ecology, Progress Series 81:269–276.
Bohannan, B. J. M., & R. E. Lenski. 1997. Effect of resource enrichment on a chemostat community of bacteria and bacteriophage. Ecology 78:2303–2315.
Boraas, M. E. 1980. A chemostat system for the study of rotifer-algal-nitrate interactions. American Society of Limnology and Oceanography, Special Symposia
3:173–182.
Borja, R., C. J. Banks, A. Martin, & B. Khalfaoui. 1995. Anaerobic digestion of
palm oil mill effluent and condensation water waste: an overall kinetic model for
methane production and substrate utilization. Bioprocess Engineering 13:87–95.
Bozdogan, H. 1987. Model selection and Akaike’s information criterion (AIC): the
general theory and its analytical extensions. Psychometrika 52:345–370.
Braumann, C. A. 1983. Population extinction probabilities and methods of estimation
for population stochastic differential equation models. In R. S. Bucy, & J. M. F.
Moura, eds., Nonlinear stochastic problems , pages 553–559. D. Reidel Publishing
Company.
Brett, M. T., & C. R. Goldman. 1997. Consumer versus resource control in freshwater
pelagic food webs. Science 275:384–386.
170
Identifying predator-prey models (PhD Thesis)
C. Jost
Bustamante, R. H., G. M. Branch, & S. Eekhout. 1995. Maintenance of an exceptional intertidal grazer biomass in south africa: subsidy by subtidal kelps. Ecology
76:2314–2329.
Cale, W. G., G. M. Henebry, & J. A. Yeakley. 1989. Inferring process from pattern
in natural communities: Can we understand what we see. BioScience 39:600–605.
Callois, J.-M. 1997. Théorie de l’interaction trophique: construction de modèles dynamiques avec interférence entre les prédateurs. Tests expérimentaux sur deux
espèces de cladocères. Dea, Institut national agronomique Paris-Grignon.
Carpenter, S. R., ed. 1988. Complex interactions in lake communities. Springer.
Carpenter, S. R., & J. F. Kitchell, eds. 1994. The Trophic Cascade in Lakes. Cambridge University Press, London.
Carpenter, S. R., J. F. Kitchell, & J. R. Hodgson. 1985. Cascading trophic interactions
and lake productivity. BioScience 35:634–639.
Carpenter, S. R., R. C. Lathrop, & A. Muñoz-del-Rio. 1993. Comparison of dynamic
models for edible phytoplankton. Canadian Journal of Fisheries and Aquatic Sciences 50:1757–1767.
Carpenter, S. R., K. L. Cottingham, & C. A. Stow. 1994. Fitting Predator-Prey
models to time series with observation errors. Ecology 75:1254–1264.
Caughley, G. 1976. Plant-herbivore systems. In R. M. May, ed., Theoretical ecology:
Principles and Applications , pages 94–113. Blackwell Scienti…c Publishers.
Characklis, W. G. 1978. Microbial reaction rate expressions. Journal of the Environmental Engineering Division 104:531–534.
Chitty, D. 1996. Do lemmings commit suicide. Oxford University Press.
Chiu, S. Y., L. E. Erickson, L. T. Fan, & I. C. Kao. 1972. Kinetic model identi…cation in mixed populations using continuous culture data. Biotechnology and
Bioengineering 14:207–231.
CIPEL. 1995. Rapports sur les études et recherches entreprises dans le bassin lémanique. Tech. rep., Commission internationale pour la protection des eaux du
Léman contre la pollution, Lausanne, Switzerland.
Clutton-Brock, M. 1967. Likelihood distributions for estimating functions when both
variables are subject to error. Technometrics 9:261–269.
Contois, D. E. 1959. Kinetics of bacterial growth: relationship between population
density and speci…c growth rate of continuous cultures. Journal of General Microbiology 21:40–50.
Cosner, C. 1996. Variability, vagueness and comparison methods for ecological models.
Bulletin of Mathematical Biology 58:207–246.
Cosner, C., D. L. DeAngelis, J. S. Ault, & D. B. Olson. in press. Effects of spatial
grouping on the functional response of predators.
Curds, C. R., & A. Cockburn. 1968. Studies on the growth and feeding of Tetrahymena pyriformis in axenic and monoxenic culture. Journal of General Microbiology
54:343–358.
C. Jost
Identifying predator-prey models (PhD Thesis)
171
Cutler, D. W., & L. M. Crump. 1924. The rate of reproduction in arti…cial culture of
Colpidium colpoda. Part III. The Biochemical Journal 18:905–912.
Daigger, G. T., & C. P. L. Grady Jr. 1977. A model for the bio-oxidation process
based on product formation concepts. Water Research 11:1049–1057.
Davidowicz, P., Z. M. Gliwicz, & R. D. Gulati. 1988. Can Daphnia prevent a bluegreen algal bloom in hypertrophic lakes? A laboratory test. Limnologica 19:21–26.
De Mazancourt, C., M. Loreau, & L. Abbadie. 1998. Grazing optimization and nutrient cycling: when do herbivores enhance plant production? Ecology 79:2242–2252.
DeAngelis, D. L., R. A. Goldstein, & R. V. O’Neill. 1975. A model for trophic interactions. Ecology 56:881–892.
Dennis, B., R. A. Desharnais, J. M. Cushing, & R. F. Costantino. 1995. Nonlinear
demographic dynamics: mathematical models, statistical methods, and biological
experiments. Ecological Monographs 65:261–281.
Dennis, B., R. A. Desharnais, J. M. Cushing, & R. F. Costantino. 1997. Transitions
in population dynamics: equilibria to periodic cycles to aperiodic cyles. Journal
of Animal Ecology 66:704–729.
Dercová, K., J. Derco, M. Hutňan, & M. Králik. 1989. Effect of formaldehyde on
kinetics of glucose consumption. Chemical papers 43:41–50.
Droop, M. R. 1966. Vitamin B12 and marine ecology. III. An experiment with a
chemostat. Journal of the Marine Biological Association of the United Kingdom
46:659–671.
Droop, M. R., & J. M. Scott. 1978. Steady-state energetics of planktonic herbivore.
Journal of the Marine Biological Association of the United Kingdom 58:749–772.
Edelstein-Keshet, L. 1988. Mathematical Models in Biology. McGraw-Hill, Inc.
Edwards, A. M., & J. Brindley. 1996. Oscillatory behaviour in a three-component
plankton population model. Dynamics and Stability of Systems 11:347–370.
Edwards, A. W. F. 1992. Likelihood. John Hopkins University, Baltimore, Maryland,
USA, expanded edn.
Efron, B., & R. J. Tibshirani. 1993. An Introduction to the Bootstrap. Chapman &
Hall.
Eigen, M., & P. Schuster. 1979. The Hypercycle. Springer-Verlag.
Ellner, S., & P. Turchin. 1995. Chaos in a noisy world: new methods and evidence
from time-series analysis. American Naturalist 145:343–375.
Ellner, S. P., B. E. Kendall, S. N. Wood, E. McCauley, & C. J. Briggs. 1997. Inferring mechanisms from time-series data: delay-differential equations. Physica D
110:182–194.
Elmaleh, S., & R. Ben Aim. 1976. In‡uence sur la cinétique biochimique de la concentration en carbone organique à l’entrée d’un réacteur développant une polyculture
microbienne en mélange parfait. Water Research 10:1005–1009.
172
Identifying predator-prey models (PhD Thesis)
C. Jost
Elton, C. 1924. Periodic ‡uctuations in the numbers of animals: their causes and
efffects. British Journal of Experimental Biology 2:119–163.
Fasham, M. J. R., H. W. Ducklow, & S. M. McKelvie. 1990. A nitrogen-based model of
plankton dynamics in the ocean mixed layer. Journal of Marine Research 48:591–
639.
Fayyaz, A. M., A. Prokop, & Z. Fencl. 1971. Growth and physiology of a yeast
cultivated in batch and continuous culture systems. Folia microbiologica 16:249–
259.
Feller, W. 1939. On the logistic law of growth and its empirical veri…cations in biology.
Acta Biotheoretica 5:51–65.
Fenchel, T. 1988. Marine Plankton Food Chains. Annual Reviews of Ecology and
Systematics 19:19–38.
Flynn, K. J., & K. Davidson. 1993. Predator-prey interactions between Isochrysis
galbana and Oxyrrhis marina. I. Changes in particulate 13 C . Journal of Plankton
Research 15:455–463.
Fott, J., B. Desortová, & J. Hrbáček. 1980. A comparison of the growth of ‡agellates
under heavy grazing stress with a continuous culture. In B. Sikyta, Z. Fencl, &
V. Poláček, eds., Continuous Cultivation of Microorganisms , vol. 7, pages 395–401,
Prague.
Fredrickson, A. G. 1977. Behavior of mixed cultures of microorganisms. Annual Review of Microbiology 31:63–87.
Free, C. A., J. R. Beddington, & J. H. Lawton. 1977. On the inadequacy of simple
models of mutual interference for parasitism and predation. Journal of Animal
Ecology 46:543–554.
Freedman, H. I., & R. M. Mathsen. 1993. Persistence in predator-prey systems with
ratio-dependent predator in‡uence. Bulletin of Mathematical Biology 55:817–827.
Fretwell, S. D. 1977. The regulation of plant communities by the food chains exploiting
them. Perspectives in Biology and Medicine 20:169–185.
Fujimoto, J. 1963. Kinetics of microbial growth and substrate consumption. Journal
of Theoretical Biology 5:171–191.
Fukaya, T., Y. Furuta, Y. Ishiguro, H. Horitsu, & K. Takamizawa. 1996. A novel model
for continuous fermentation process for Worcestershire sauce production using a
trickle bed bioreactor. Journal of Fermentation and Bioengineering 81:233–239.
Gatto, M. 1991. Some Remarks on models of plankton densities in lakes. American
Naturalist 137:264–267.
Gatto, M. 1993. The evolutionary optimality of oscillatory and chaotic dynamics in
simple population models. Theoretical Population Biology 43:310–336.
Gause, G. F. 1934. The Struggle for Existence. Hafner Publishing Company, Inc.,
New York.
Gause, G. F. 1935. Experimental demonstrations of Volterra’s periodic oscillations in
the numbers of animals. British Journal of Experimental Biology 12:44–48.
C. Jost
Identifying predator-prey models (PhD Thesis)
173
Gause, G. F., N. P. Smaragdova, & A. A. Witt. 1936. Further studies of interactions
between predators and prey. Journal of Animal Ecology 5:1–18.
Gawler, M., G. Balvay, P. Blanc, J.-C. Druart, & J. P. Pelletier. 1988. Plankton ecology of Lake Geneva: A test of the PEG-model. Archiv für Hydrobiologie 114:161–
174.
Getz, W. M. 1984. Population dynamics: a per capita resource approach. Journal of
Theoretical Biology 180:623–643.
Ghaly, A. E., & E. A. Echiegu. 1993. Kinetics of a continuous-‡ow no-mix anaerobic
reactor. Energy sources 15:433–449.
Gilpin, M., & F. J. Ayala. 1973. Global models of growth and competition. Proceedings
of the National Academy of Sciences of the USA 70:3590–3593.
Gilpin, M., T. J. Case, & F. J. Ayala. 1976. Theta-selection. Mathematical Biosciences
32:131–139.
Ginzburg, L. R. 1998. Assuming reproduction to be a function of consumption raises
doubts about some popular predator-prey models. Journal of Animal Ecology
67:325–327.
Ginzburg, L. R., & H. R. Akçakaya. 1992. Consequences of ratio-dependent predation
for steady-state properties of ecosystems. Ecology 73:1536–1543.
Gleeson, S. K. 1994. Density dependence is better than ratio dependence. Ecology
75:1834–1835.
Gliwicz, Z. M. 1990. Why do cladocerans fail to control algal blooms? Hydrobiologia
200/201:83–97.
Goel, N. S., S. C. Maitra, & E. W. Montroll. 1971. On the Volterra and other nonlinear
models of interacting populations. Reviews of Modern Physics 43:231–276.
Goldman, J. C. 1977. Steady state growth of phytoplankton in continuous culture:
Comparison of internal and external nutrient equations. Journal of Phycology
13:251–258.
Goma, G., & D. Ribot. 1978. Hydrocarbon fermentation: Kinetics of microbial cell
growth. Biotechnology and Bioengineering 20:1723–1734.
Gomatam, J. 1974. A new model for interacting populations. I. Two-species systems.
Bulletin of Mathematical Biology 36:347–353.
Gompertz, B. 1825. On the nature of the function expressive of the law of human
mortality. Philosophical Transactions 115:513–585.
Gouriéroux, C., & A. Monfort. 1996. Simulation-based econometric methods. Oxford
University Press.
Grady Jr., C. P. L., & D. R. Williams. 1975. Effects of in‡uent substrate concentration
on the kinetics of natural microbial populations in continuous culture. Water
Research 9:171–180.
Grady Jr., C. P. L., L. J. Harlow, & R. R. Riesing. 1972. Effects of growth rate and
in‡uent substrate concentration on effluent quality from chemostats containing
174
Identifying predator-prey models (PhD Thesis)
C. Jost
bacteria in pure and mixed culture. Biotechnology and Bioengineering 14:391–
410.
Gragnani, A. 1997. Bifurcation analysis of two predator-prey models. Applied Mathematics and Computation 85:97–108.
Gutierrez, A. 1992. The physiological basis of ratio-dependent predator-prey theory:
a metabolic pool model of Nicholson’s blow‡ies as an example. Ecology 73:1552–
1563.
Hairston, N. G., F. E. Smith, & L. B. Slobodkin. 1960. Community structure, population control, and competition. American Naturalist 44:421–425.
Hanski, I., & E. Korpimäki. 1995. Microtine rodent dynamics in northern europe:
parameterized models for the predator-prey interaction. Ecology 76:840–850.
Hansson, L.-A., P. N. Brönmark, L. Greenberg, P. Lundberg, P. A. Nilsson, A. Persson, L. B. Pettersson, P. Romare, & L. J. Tranvik. 1998. Consumption patterns,
complexity and enrichment in aquatic food chains. Proceedings of the Royal Society
of London, B 265:901–906.
Harris, G. P. 1996. A reply to Sarnelle (1996) and some further comments on Harris’s
(1994) opinions. Freshwater Biology 35:343–347.
Harrison, G. W. 1995. Comparing predator-prey models to Luckinbill’s experiment
with Didinium and Paramecium. Ecology 76:157–174.
Hassell, M. P. 1978a. The dynamics of arthropod predator-prey systems. Princeton
University Press.
Hassell, M. P. 1978b. Sigmoid functional responses and population stability. Theoretical Population Biology 14:62–67.
Hassell, M. P., & D. J. Rogers. 1972. Insect parasite responses in the development of
population models. Journal of Animal Ecology 41:661–676.
Hassell, M. P., & G. C. Varley. 1969. New inductive population model for insect
parasites and its bearing on biological control. Nature 223:1133–1137.
Hassell, M. P., J. H. Lawton, & J. R. Beddington. 1977. Sigmoid functional response
by invertebrate predators and parasitoids. Journal of Animal Ecology 46:249–262.
Hastings, A. 1997. Population biology: concepts and models. Springer.
Herbert, D., R. Elsworth, & R. C. Telling. 1956. The continuous culture of bacteria: a
theoretical and experimental study. Journal of General Microbiology 14:601–622.
Hilborn, R., & M. Mangel. 1997. The Ecological Detective. Princeton University
Press.
Hilborn, R., & C. J. Walters. 1992. Quantitative Fisheries Stock Assessment. Chapman and Hall.
Holling, C. S. 1959a. The components of predation as revealed by a study of smallmammal predation of the European pine saw‡y. The Canadian Entomologist
91:293–320.
C. Jost
Identifying predator-prey models (PhD Thesis)
175
Holling, C. S. 1959b. Some characteristics of simple types of predation and parasitism.
The Canadian Entomologist 91:385–398.
Holmgren, N., S. Lundberg, & P. Yodzis. 1996. Functional responses and heterogeneities: a reanalysis of an experiment with cladocerans. Oikos 76:196–198.
Hosten, L. H. 1974. A sequential experimental design procedure for precise parameter estimation based upon the shape of the joint con…dence region. Chemical
Engineering Science 29:2247–2252.
Hsu, S.-B., & T.-W. Huang. 1995. Global stability for a class of predator-prey systems.
SIAM Journal of Applied Mathematics 55:763–783.
Huet, S., E. Jolivet, & A. Messéan. 1992. La régression non-linéaire: méthodes et
applications en biologie. INRA Editions, Paris.
Huffaker, C. B. 1958. Experimental studies on predation: dispersion factors and
predator-prey oscillations. Hilgardia 27:343–383.
Huffaker, C. B., K. P. Shea, & S. G. Herman. 1963. Experimental studies on predation:
Complex dispersion and levels of food in an acarine predator-prey interaction.
Hilgardia 34:303–330.
Hunter, M. D., & P. W. Price. 1992. Playing chutes and ladders: heterogeneity and the
relative roles of bottom-up and top-down forces in natural communities. Ecology
73:724–732.
Ivlev, V. S. 1961. Experimental Ecology of the Feeding of Fishes. Yale University
Press, New Haven, CT.
Jannasch, H. W. 1967. Growth of marine bacteria at limiting concentrations of organic
carbon in seawater. Limnology and Oceanography 12:264–271.
Jannasch, H. W., & T. Egli. 1993. Microbial growth kinetics: a historical perspective.
Antonie van Leeuwenhoek 63:213–224.
Jeppesen, E., M. Søndergaard, O. Sortkjaer, E. Mortensen, & P. Kristensen. 1990.
Interactions between phytoplankton, zooplankton and …sh in a shallow hypertrophic lake: a study of phytoplankton collapses in Lake Søbygård, Denmark.
Hydrobiologia 191:149–164.
Jones, R. H. 1993. Longitudinal data with serial correlation: a state-space approach,
vol. 47 of Monographs on Statistics and Applied Probability . Chapman and Hall,
London.
Jost, C., O. Arino, & R. Arditi. 1999. About deterministic extinction in ratiodependent predator-prey models. Bulletin of Mathematical Biology 61:19–32.
Jost, J. L., J. F. Drake, H. M. Tsuchiya, & A. G. Fredrickson. 1973. Microbial food
chains and food webs. Journal of Theoretical Biology 41:461–484.
Kantz, H., & T. Schreiber. 1997. Nonlinear time series analysis. Cambridge University
Press.
Kargi, F., & M. L. Shuler. 1979. Generalized differential speci…c rate equation for
microbial growth. Biotechnology and Bioengineering 21:1871–1875.
176
Identifying predator-prey models (PhD Thesis)
C. Jost
Kaunzinger, C. M. K., & P. J. Morin. 1998. Productivity controls food chain properties in microbial communities. Nature 383:in press.
Krebs, J. R., & N. B. Davies. 1993. An introduction to behavioural ecology. Blackwell,
3rd edn.
Kristiansen, B., & C. G. Sinclair. 1979. Production of citric acid in continuous culture.
Biotechnology and Bioengineering 21:297–315.
Kuang, Y., & E. Beretta. 1998. Global qualitative analysis of a ratio-dependent
predator-prey system. Journal of Mathematical Biology 36:389–406.
Lampert, W. 1978. Climatic conditions and planctonic interactions as factors controlling the regular succession of spring algal bloom and extremely clear water in
Lake Constance. Verhandlungen der Internationalen Vereinigung für Limnologie
20:969–974.
Lampert, W. 1985. The role of zooplankton: An attempt to quantity grazing. In Lakes
Pollution and Recovery , pages 54–62, Rome. Proceedings of the International
Congress of the European Water Pollution Control Association.
Lampert, W., W. Fleckner, H. Rai, & B. E. Taylor. 1986. Phytoplankton control
by grazing zooplankton: a study on the spring clear water phase. Limnology and
Oceanography 31:478–490.
Lavigne, D. M. 1996. Ecological interactions between marine mammals, commercial
…sheries, and their prey: unravelling the tangled web. In W. A. Montevecchi, ed.,
Studies of high-latitude seabirds. 4. Trophic relationships and energetics of endotherms in cold ocean systems , pages 59–71. Canadian Wildlife Service, Ottawa,
Canada.
Lek, S., M. Delacoste, P. Baran, I. Dimopoulos, j. Lauga, & S. Aulagnier. 1996.
Application of neural networks to modelling nonlinear relationships in ecology.
Ecological Modelling 90:39–52.
Lequerica, J. L., S. Vallés, & A. Flors. 1984. Kinetics of rice straw fermentation.
Applied Microbiology and Biotechnology 19:70–74.
Leslie, P. H. 1948. Some further notes on the use of matrices in population mathematics. Biometrika 35:213–245.
Leslie, P. H., & J. C. Gower. 1960. The properties of a stochastic model for the
predator-prey type of interaction between two species. Biometrika 47:219–234.
Linhard, H., & W. Zucchini. 1986. Model Selection. John Wiley & Sons.
Lotka, A. J. 1924. Elements of Physical Biology. Williams and Wilkins Co.
Luck, R. F. 1990. Evaluation of natural enemies for biological control: a behavioral
approach. Trends in Ecology and Evolution 5:196–199.
Luckinbill, L. S. 1973. Coexistence in laboratory populations of Paramecium aurelia
and its predator Didinium nasutum. Ecology 54:1320–1327.
Lundberg, P., & J. M. Fryxell. 1995. Expected population density versus productivity
in ratio-dependent and prey-dependent models. American Naturalist 146:153–161.
C. Jost
Identifying predator-prey models (PhD Thesis)
177
Lynch, M., J. Conery, & R. Bürger. 1995. Mutation accumulation and the extinction
of small populations. American Naturalist 146:489–518.
Malthus, T. R. 1798. An Essay on the Principle of Population, and A Summary View
of the Principle of Population. Penguin, Harmondsworth, England.
Mansour, R. A., & R. N. Lipcius. 1991. Density-dependent foraging and mutual
interference in blue crabs preying upong infaunal clams. Marine Ecology Progress
Series 72:239–246.
May, R. M. 1973. Stability and Complexity in Model Ecosystems. Princeton University Press, Princeton.
May, R. M. 1975. Stability and Complexity in Model Ecosystems. Princeton University Press, Princeton, 2nd edn.
May, R. M. 1976a. Models for single populations. In R. M. May, ed., Theoretical
ecology: Principles and Applications , pages 4–25. Blackwell Scienti…c Publishers.
May, R. M. 1976b. Models for two interacting populations. In R. M. May, ed., Theoretical ecology: Principles and Applications , pages 49–70. Blackwell Scienti…c
Publishers.
May, R. M. 1989. Detecting density dependence in imaginary worlds. Nature 338:16–
17.
Maynard Smith, J. 1975. Models in Ecology. Cambridge University Press.
Mazumder, A. 1994. Patterns of algal biomass in dominant odd- vs. even-link lake
ecosystems. Ecology 75:1141–1149.
Mazumder, A., & D. R. S. Lean. 1994. Consumer-dependent responses of lake ecosystems to nutrient loading. Journal of Plankton Research 16:1567–1580.
McCarthy, M. A., L. R. Ginzburg, & H. R. Akçakaya. 1995. Predator interference
across trophic chains. Ecology 76:1310–1319.
McCauley, E., & W. W. Murdoch. 1987. Cyclic and stable populations: plankton as
a paradigm. American Naturalist 129:97–121.
McCauley, E., W. Murdoch, & S. Watson. 1988. Simple models and variation in
plankton densities among lakes. American Naturalist 132:383–403.
Menge, B. A. 1992. Community regulation: under what conditions are bottom-up
factors important on rocky shores? Ecology 73:755–765.
Michalski, J., & R. Arditi. 1995a. Food web structure at equilibrium and far from it:
is it the same? Proceedings of the Royal Society of London, B 259:217–222.
Michalski, J., & R. Arditi. 1995b. Food webs with predator interference. Journal of
biological systems 3:323–330.
Michalski, J., J.-C. Poggiale, R. Arditi, & P. M. Auger. 1997. Macroscopic dynamic
effects of migrations in patchy predator-prey systems. Journal of Theoretical Biology 185:459–474.
Mittelbach, G. G., C. W. Osenberg, & M. A. Leibold. 1988. Trophic relations and
ontogenetic niche shifts in aquatic ecosystems. In B. Ebenmann, & L. Persson,
eds., Size-structured populations , pages 219–235. Springer.
178
Identifying predator-prey models (PhD Thesis)
C. Jost
Monod, J. 1942. Recherches sur la croissance des cultures bactérienne. Hermann et
Cie, Paris.
Moore, M. V. 1988. Density-dependent predation of early instar Chaoborus feeding
on multispecies prey assemblages. Limnology and Oceanography 33:256–268.
Morris, W. F. 1997. Disentangling effects of induced plant defenses and food quantity
on herbivores by …tting nonlinear models. American Naturalist 150:299–327.
Morrison, K. A., N. Thérien, & B. Marcos. 1987. Comparison of six models for nutrient
limitations on phytoplankton growth. Canadian Journal of Fisheries and Aquatic
Sciences 44:1278–1288.
Moser, A. 1958. The dynamics of bacterial populations maintained in the chemostat.
Ph.D. thesis, The Carnegie Institute, Washington.
Murdoch, W. A., & A. Oaten. 1975. Predation and population stability. Advances in
Ecological Research 9:2–132.
Murdoch, W. W., J. Chesson, & P. L. Chesson. 1985. Biological control in theory and
practice. American Naturalist 125:344–366.
Murdoch, W. W., R. M. Nisbet, E. McCauley, A. M. DeRoos, & W. S. C. Gurney.
1998. Plankton abundance and dynamics across nutrient levels: tests of hypothesis. Ecology 79:1339–1356.
Myerscough, M. R., M. J. Darwen, & W. L. Hogarth. 1996. Stability, persistence
and structural stability in a classical predator-prey model. Ecological Modelling
89:31–42.
Nicholson, A. J. 1933. The balance of animal populations. Journal of Animal Ecology
2:131–178.
Nisbet, R. M., & W. S. C. Gurney. 1982. Modelling Fluctuating Populations. John
Wiley & Sons.
Oksanen, L. 1983. Trophic exploitation and arctic phytomass patterns. American
Naturalist 122:45–52.
Oksanen, L., S. D. Fretwell, J. Arruda, & P. Niemelä. 1981. Exploitation ecosystems
in gradients of primary productivity. American Naturalist 118:240–261.
Pareilleux, A., & N. Chaubet. 1980. Growth kinetics of apple plant cell cultures.
Biotechnology Letters 2:291–296.
Pascual, M. A., & P. Kareiva. 1996. Predicting the outcome of competition using
experimental data: maximum likelihood and bayesian approaches. Ecology 77:337–
349.
Pavé, A. 1994. Modélisation en biologie et en écologie. Aléas Editeur, Lyon.
Persson, L., G. Andersson, S. F. Hamrin, & L. Johansson. 1988. Predator regulation and primary productivity along the productivity gradient of temperate lake
ecosystems. In S. R. Carpenter, ed., Complex Interactions in Lake Communities ,
pages 45–65. Springer Verlag, New York.
Philip, J. R. 1957. Sociality and sparse populations. Ecology 38:107–111.
C. Jost
Identifying predator-prey models (PhD Thesis)
179
Pimm, S. L. 1982. Food Webs. Chapman and Hall, London.
Pimm, S. L. 1991. The Balance of Nature. University of Chicago Press.
Poggiale, J.-C., J. Michalski, & R. Arditi. 1998. Emergence of donor control in patchy
predator-prey systems. Bulletin of Mathematical Biology 60:in press.
Ponsard, S. 1998. Trophic structure (isotopic study) and regulation of abundances in
the forest litter macrofauna. Ph.D. thesis, Institut national agronomique, ParisGrignon.
Porter, K. G. 1977. The plant-animal interface in freshwater ecosystems. American
Scientist 65:159–170.
Possingham, H. P., S. Tuljapurkar, J. Roughgarden, & M. Wilks. 1994. Population
cycling in space-limited organisms subject to density-dependent predation. American Naturalist 143:563–582.
Power, M. E. 1992. Top-down and bottom-up forces in food webs: do plants have
primacy? Ecology 73:733–746.
Press, W. H., S. A. Teukolsky, W. T. Vetterling, & B. P. Flannery. 1992. Numerical Recipes in C: The Art of Scienti…c Computing. Cambridge University Press,
Cambridge, 2nd edn.
Ratkowsky, D. A. 1983. Nonlinear Regression Modeling. Marcel Dekker, Inc.
Real, L. A. 1977. The kinetics of the functional response. American Naturalist
111:289–300.
Reilly, P. M., & H. Patino-Leal. 1981. A bayesian study of the error-in-variables
model. Technometrics 23:221–231.
Richards, F. J. 1959. A ‡exible growth function for empirical use. Journal of experimental Botany 29:290–300.
Rogers, D. 1972. Random search and insect population models. Journal of Animal
Ecology 41:369–383.
Rogers, D., & M. P. Hassell. 1974. General models for insect parasite and predator
searching behaviour: interference. Journal of Animal Ecology 43:239–253.
Roques, H., S. Yue, S. Saipanich, & B. Capdeville. 1982. Faut-il abandonner le formalisme de monod pour la modélisation des processus de dépollution par voie
biologique ? Water Research 16:839–847.
Rosenzweig, M. L. 1969. Why the prey curve has a hump. American Naturalist
103:81–87.
Rosenzweig, M. L. 1971. Paradox of enrichment: destabilization of exploitation
ecosystems in ecological time. Science 171:385–387.
Rosenzweig, M. L., & R. H. MacArthur. 1963. Graphical representation and stability
conditions of predator-prey interactions. American Naturalist 97:217–223.
Royama, T. 1971. A comparative study of models for predation and parasitism. Researches on Population Ecology (Kyoto) Supplement 1:1–91.
180
Identifying predator-prey models (PhD Thesis)
C. Jost
Rubel, L. A. 1981. A universal differential equation. Bulletin (new series) of the
american mathematical society 4:345–349.
Ruxton, G. D. 1995. Short term refuge use and stability of predator-prey models.
Theoretical Population Biology 47:1–17.
Ruxton, G. D., & W. S. C. Gurney. 1992. The interpretation of tests for ratiodependence. Oikos 65:334–335.
Ruxton, G. D., & W. S. C. Gurney. 1994. Deriving the functional response without
assuming homogeneity. American Naturalist 144:537–541.
Ruxton, G. D., W. S. C. Gurney, & A. M. de Roos. 1992. Interference and generation
cycles. Theoretical Population Biology 42:235–253.
Salt, G. W. 1974. Predator and prey densities as controls of the rate of capture by
the predator Didinium nasutum. Ecology 55:434–439.
Sarnelle, O. 1992. Nutrient enrichment and grazer effects on phytoplankton in lakes.
Ecology 73:551–560.
Sarnelle, O. 1994. Inferring process from pattern: trophic level abundances and imbedded interactions. Ecology 75:1835–1841.
Sarnelle, O., S. D. Cooper, S. Wiseman, & K. Mavuti. 1998. The relationship between
nutrients and trophic-level biomass in turbid tropical ponds. Freshwater Biology
40:65–75.
Sas, H. 1989. Lake Restoration by Reduction of Nutrient Loading: Expectations,
Experiences, Extrapolations. Academia-Verlag Richarz, Sankt Augustin.
Scheffer, M. 1998. Ecology of Shallow Lakes. Chapman & Hall.
Scheffer, M., & R. J. de Boer. 1995. Implications of spatial heterogeneity for the
paradox of enrichment. Ecology 76:2270–2277.
Schnute, J. T. 1994. A general framework for developing sequential …sheries models.
Canadian Journal of Fisheries and Aquatic Sciences 51:1676–1688.
Schnute, J. T., & L. J. Richards. 1995. The in‡uence of error on population estimates from catch-age models. Canadian Journal of Fisheries and Aquatic Sciences
52:2063–2077.
Schoener, T. W. 1978. Effects of density-restricted food encounter on some single-level
competition models. Theoretical Population Biology 13:365–381.
Seber, G. A. F., & C. J. Wild. 1989. Nonlinear Regression. Wiley.
Seda, J. 1989. Main factors affecting spring development of herbivorous Cladocera in
the Římov Reservoir (Czechoslovakia). Archiv für Hydrobiologie. Beihefte: Ergebnisse der Limnologie 33:619–630.
Shea, K. 1998. Management of populations in conservation, harvesting and control.
Trends in Ecology and Evolution 13:371–375.
Sherr, E. B., & B. F. Sherr. 1991. Planktonic microbes: tiny cells at the base of the
ocean’s food webs. TREE 6:50–54.
C. Jost
Identifying predator-prey models (PhD Thesis)
181
Sherratt, J. A., B. T. Eagan, & M. A. Lewis. 1997. Oscillations and chaos behind
predator-prey invasion: mathematical artifact or ecological reality? Philosophical
Transactions of the Royal Society of London, B 352:21–38.
Sih, A. 1984. Optimal behaviour and density-dependent predation. American Naturalist 123:314–236.
Simons, T. J., & D. C. L. Lam. 1980. Some limitations of water quality models for
large lakes: a case study of lake Ontario. Water Resources Research 16:105–116.
Smith, F. E. 1963. Population dynamics in Daphnia magna and a new model for
population growth. Ecology 44:651–663.
Sokol, W., & J. A. Howell. 1981. Kinetics of phenol oxidation by washed cells. Biotechnology and Bioengineering 23:2039–2049.
Solomon, M. E. 1949. The natural control of animal populations. Journal of Animal
Ecology 18:1–35.
Solow, A. R. 1995. Fitting population models to time series data. In T. M. Powell, &
J. H. Stelle, eds., Ecological time series . Chapman & Hall.
Sommer, U., Z. M. Gliwicz, W. Lampert, & A. Duncan. 1986. The PEG-Model of
seasonal succession of planktonic events in fresh waters. Archiv für Hydrobiologie
106:433–471.
Steele, J., & E. Henderson. 1992. The role of predation in plankton models. Journal
of Plankton Research 14:157–172.
Steele, J. H., & E. W. Henderson. 1981. A simple plankton model. American Naturalist 117:676–691.
Stone, L., & T. Berman. 1993. Positive feedback in aquatic ecosystems: the case of
the microbial loop. Bulletin of Mathematical Biology 55:919–936.
Stow, C. A., S. R. Carpenter, & K. L. Cottingham. 1995. Resource versus ratiodependent consumer-resource models: a Bayesian perspective. Ecology 76:1986–
1990.
Strebel, D. E., & N. S. Goel. 1973. On the isocline methods for analyzing preypredator interactions. Journal of Theoretical Biology 39:211–234.
Strong, D. R. 1992. Are trophic cascades all wet? Differentiation and donor-control
in speciose ecosystems. Ecology 73:747–754.
Sutherland, W. J. 1983. Aggregation and the ’ideal free’ distribution. Journal of
Animal Ecology 52:821–828.
Sutherland, W. J. 1996. From individual behaviour to population ecology. Oxford
University Press.
Svirezhev, Y. M., & D. O. Logofet. 1983. Stability of Biological Communities (English
ed.). Mir Publishers, Moscow.
Szathmáry, E. 1991. Simple growth laws and selection consequences. Trends in Ecology and Evolution 6:366–370.
182
Identifying predator-prey models (PhD Thesis)
C. Jost
Takahashi, F. 1964. Reproduction curve with two equilibrium points: a consideration
on the ‡uctuation of insect populations. Researches on Population Ecology (Kyoto)
6:28–36.
Tanner, J. T. 1975. The stability and the intrinsic growth rates of prey and predator
populations. Ecology 56:855–867.
Taub, F. B., & D. H. McKenzie. 1973. Continuous cultures of an alga and its grazer.
Bulletins from the Ecological Research Committee 17:371–377.
Teissier, G. 1936. Quantitative laws of growth. Annales de Physiologie et Physiochimie
Biologique 12:527–586.
Thompson, W. R. 1924. La théorie mathématique de l’action des parasites entomophages et le facteur du hasard. Annales de la Faculté des Sciences Marseille
2:69–89.
Tijero, J., E. Guardiola, M. Cortija, & L. Moreno. 1989. Kinetic study of anaerobic
digestion of glucose and sucrose. Journal of Environmental Science and Health
A24:297–319.
Tostowaryk, W. 1972. The effect of prey defence on the functional response of Podisus
modestus (Hemiptera: Pentatomidae) to densities of the saw‡ies Neodiprion
swainei and N. pratti banksianae (Hymenoptera: Neodiprionidae). The Canadian
Entomologist 104:61–69.
Trexler, J. C. 1988. How can the functional response best be determined? Oecologia
76:206–214.
Turchin, P., & S. Ellner. (unp.). Living on the edge of chaos: population dynamics of
Fennoscandian voles.
Turchin, P., & I. Hanski. 1997. An empirically based model for latitudinal gradient
vole population dynamics. American Naturalist 149:842–874.
Turchin, P., & A. D. Taylor. 1992. Complex dynamics in ecological time series. Ecology
73:289–305.
Utida, S. 1950. On the equilibrium state of the interacting population of an insect
and its parasite. Ecology 31:165–175.
van der Meer, J., & B. J. Ens. 1997. Models of interference and their consequences for
the spatial distribution of ideal free predators. Journal of Animal Ecology 66:846–
858.
Varley, G. C. 1947. The natural control of population balance in the knapweed gall-‡y
(Urophora jaceana). Journal of Animal Ecology 16:139–187.
Veilleux, B. G. 1979. An analysis of the predatory interaction between Paramecium
and Didinium. Journal of Animal Ecology 48:787–803.
Verhulst, P. F. 1838. Notice sur la loi que la population suit dans son accroissement.
Correspondances Mathématiques et physiques 10:113–121.
Volterra, V. 1926. Fluctuations in the abundance of a species considered mathematically. Nature 118:558–560.
C. Jost
Identifying predator-prey models (PhD Thesis)
183
Volterra, V. 1938. Population growth, equilibria, and extinction under speci…ed breeding conditions: a development and extension of the theory of the logistic curve.
Human Biology 10:1–11.
von Bertalanffy, L. 1951. Theoretische Biologie, vol. 2. A. Frank, Bern, 2nd edn.
Vyhnálek, V., J. Komárková, J. Seda, Z. Brandl, K. Šimek, & N. Johanisová. 1991.
Clear-water phase in the Římov Reservoir (South Bohemia): Controlling factors.
Verhandlungen der Internationalen Vereinigung für Limnologie 24:1336–1339.
Vyhnálek, V., J. Seda, & J. Nedoma. 1993. Fate of the spring phytoplankton bloom
in Římov Reservoir (Czechoslovakia): Grazing, lysis and sedimentation. Verhandlungen der Internationalen Vereinigung für Limnologie 25:1192–1195.
Vyhnálek, V., J. Hejzlar, J. Nedoma, & J. Vrba. 1994. Importance of the river in‡ow for the spring development of plankton in Římov Reservoir (Czechoslovakia).
Verhandlungen der Internationalen Vereinigung für Limnologie 40:51–56.
Walter, E., ed. 1987. Identi…ability of Parametric Models. Pergamon Press.
Walter, E., & L. Pronzato. 1997. Identi…cation of Parametric Models from Experimental Data. Springer.
Watt, K. E. F. 1959. A mathematical model for the effect of densities of attacked and
attacking species on the number attacked. The Canadian Entomologist 91:129–
144.
Weisse, T. H. M., R. M. Pinto-Coelho, A. Schweizer, D. Springmann, & G. Baldringer.
1990. Response of the microbial loop to the phytoplankton spring bloom in a large
prealpine lake. Limnology and Oceanography 35:781–794.
Wilhelm, R. 1993. Dynamics and persistence of a microbial predator-prey system in
laboratory culture. Tech. Rep. 2835, ISSN 0944-2952, Forschungszentrum Jülich,
Arbeitsgruppe theoretische Oekologie, Zentralbibliothek, D-52425 Jülich.
Yodzis, P. 1989. Introduction to Theoretical Ecology. Harper & Row.
Yodzis, P. 1994. Predator-prey theory and management of multispecies …sheries. Ecological Applications 4:51–58.
Yodzis, P. 1995. Food webs and perturbation experiments: theory and practice. In
G. A. Polis, & K. O. Winemiller, eds., Food Webs: Integrations of Patterns and
Dynamics , pages 192–200. Chapman and Hall.
Zheng, D. W., J. Bengtsson, & I. Å. Göran. 1997. Soil food webs and ecosystem
processes: decomposition in donor-control and Lotka-Volterra systms. American
Naturalist 149:125–148.
Zielinski, W. S., & L. E. Orgel. 1987. Autocatalytic synthesis of a tetranucleotide
analogue. Nature 327:346–347.
♥
Titel: En qualitative und quantitative Vergliich zwüschet Räuber-Beute
Model und ökologische Ziitreihene
Zämefassig
Di vorligendi Dissertation vergliicht zwei Räuber-Beute Model mit Ziitreihene, entweder us em
Labor oder vo Feld-Studie. S’erschti Model nimmt a dass d’Konsumfunktion nume vo de Beute
Dichti abhanget. So n’es Model zeiget charakteristike vo n’ere vo-obe-duraab Kontrolle. S’anderi Model nimmt a dass d’Konsumfunktion vom Verhältnis Beute pro Räuber abhanget. Die
Ahnahm hätt zur Folg dass das Model nöd nume vo-obe-duraab Kontrolle zeiget, sondern au
vo-une-duruuf Kontrolle. Di mathematisch Analyse vo dem Verhältnis-abhängige Model zeigt e
richi Grenz-Dünamik mit mehrere Attraktore. Eine vo dene Attraktore isch de Ursprung (das
heisst beidi populatione stärbed us). De hauptsächlich Unterschiid zwüsched dene zwei Model
isch ihri Reaktion zu n’ere Beriicherig vo de Umwelt: s’erschti Model verlürt sini Stabilität und
nume de Räuber pro…tiert vom Riichtum, wohingege s’zweiti Model gliich stabil bliibt und beidi,
de Räuber und d’Beute, pro…tiered vom Riichtum. De Vergliich vo dene beide Model mit em
verbale PEG-Model (beobachteti Dünamik vo Plankton i Süesswasser See i de gmässigte KlimaZone) zeiget dass beidi Model die Dünamik chönd erkläre falls sich ein oder zwei vo de Model
Parameter mit de Saison verändered. Di maximale Ähnlichkeit isch s’statistische Werkzüg zum
die beide Model quantitativ mit Ziitreihene vo verschidene Tüpe vo Räuber-Beute Syschtem
z’vergliiche. Einzeller- und Insekte-Date (beidi us em Labor) passed generell besser zum erschte
(nume Beute-abhängige) Model, wohingege d’Plankton Date mit beidn’e Model übereinstimmed
ohni ei’s vo de Model z’bevorzüche. I dem Fall isch es nützlich beidi Model z’Bruuche, erschtens, zum use…nde welli Vorhersage emp…ndlich uf de Underschiid zwüsched dene zwei Model
reagiered, und zweitens, zum wiiteri Forschige inschpiriere.
Title: Comparing predator-prey models qualitatively and quantitatively
with ecological time-series data
Abstract
This thesis compares two predator-prey models with the temporal dynamics observed in laboratory or …eld predator-prey systems. The …rst model assumes that the functional response is
a function of prey density only, resulting in a model with top-down characteristics. In contrast,
the second model considers this function to depend on the ratio prey per predator, resulting in
a model with both top-down and bottom-up characteristics. The mathematical analysis of this
ratio-dependent model reveals rich boundary dynamics with multiple attractors, one of them
being the origin (extinction of both populations). The major difference between the two models
is in the predictions in response to enrichment, which acts as a destabilizing factor and increass
the predator equilibrium only in the prey-dependent model, but which is neutral with respect
to stability and increases both prey and predator equilibria in the ratio-dependent model. The
comparison of both models to the verbal PEG-model (observed plankton dynamics in freshwater lakes) shows that they both can explain these dynamics if seasonality is added to one or
several model parameters. The maximum likelihood concept is used to compare the two models
quantitatively with predator-prey time series of several types. Protozoan and arthropod (laboratory) data are generally better described by the prey-dependent model. For the phytoplanktonzooplankton interaction, both models are valid and none is better than the other. In this case,
using both models can help detecting predictions that are sensitive to predator dependence and
direct further research if necessary.
Discipline : Ecology
Keywords : predator-prey models, predator-prey dynamics, nonlinear regression, model-selection, time series, ratio dependence, PEG-model, Contois model
Titre: Comparaison qualitative et quantitative de modèles proie-prédateur
à des données chronologiques
Résumé
La présente thèse compare deux modèles proie-prédateur avec les dynamiques temporelles de
systèmes observés en laboratoire ou sur le terrain. Le premier modèle suppose que la réponse
fonctionnelle dépend uniquement de la densité des proies, et présente donc les caractéristiques
des modèles où les abondances sont contrôlées “de haut en bas”. Au contraire, le second modèle
considère que la réponse fonctionnelle dépend du ratio entre densité de proies et densité de prédateurs, et inclut donc une régulation des abondances “de bas en haut”. L’analyse mathématique
de ce modèle ratio-dépendant fait apparaître des dynamiques de bord riches avec de multiples
attracteurs, dont l’un est l’origine (extinction des deux populations). La différence majeure
entre les deux modèles réside dans leurs prédictions sur la réponse d’un système à l’enrichissement: déstabilisation, et augmentation de l’abondance à l’équilibre du prédateur uniquement
dans le modèle proie-dépendant, stabilité inchangée et augmentation de l’abondance à l’équilibre des proies et des prédateurs dans le modèle ratio-dépendant. La comparaison de ces deux
modèles avec le modèle verbal PEG (décrivant la dynamique planctonique dans les lacs) montre que tous deux peuvent rendre compte de cette dynamique si des changements saisonniers
sont introduits dans les valeurs d’un ou plusieurs paramètres. Nous comparons quantitativement les deux modèles avec différents types de séries temporelles de systèmes proie-prédateur
par la méthode du maximum de vraisemblance. Les données concernant des protozoaires ou
des arthropodes (en laboratoire) sont en général mieux décrites par le modèle proie-dépendant.
Pour l’interaction phytoplancton-zooplancton, les deux modèles conviennent aussi bien l’un que
l’autre. Le fait d’utiliser les deux modèles peut alors permettre de détecter parmi les prédictions
celles qui sont sensibles à la prédateur-dépendance et, éventuellement, d’orienter des recherches
supplémentaires.
Title: Comparing predator-prey models qualitatively and quantitatively
with ecological time-series data
Abstract: see preceeding page
Discipline : Ecologie
Mots-Clés : modèles proie-prédateur, dynamiques proie-prŐdateur, regression nonlinéaire, sélection de modèle, séries chronologiques, ratio-dépendance,
modèle de Contois, modèle PEG
Ecologie de populations et communautés, URA-CNRS 2154, Bât. 362, Université ParisSud XI, 91405 Orsay cedex