Skip to main content
Log in

An updated perspective on the role of environmental autocorrelation in animal populations

  • ORIGINAL PAPER
  • Published:
Theoretical Ecology Aims and scope Submit manuscript

Abstract

Ecological theory predicts that the presence of temporal autocorrelation in environments can considerably affect population extinction risk. However, empirical estimates of autocorrelation values in animal populations have not decoupled intrinsic growth and density feedback processes from environmental autocorrelation. In this study, we first discuss how the autocorrelation present in environmental covariates can be reduced through nonlinear interactions or by interactions with multiple limiting resources. We then estimated the degree of environmental autocorrelation present in the Global Population Dynamics Database using a robust, model-based approach. Our empirical results indicate that time series of animal populations are affected by low levels of environmental autocorrelation, a result consistent with predictions from our theoretical models. Claims supporting the importance of autocorrelated environments have been largely based on indirect empirical measures and theoretical models seldom anchored in realistic assumptions. It is likely that a more nuanced understanding of the effects of autocorrelated environments is necessary to reconcile our conclusions with previous theory. We anticipate that our findings and other recent results will lead to improvements in understanding how to incorporate fluctuating environments into population risk assessments.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5

Similar content being viewed by others

References

  • Abbott KC, Ripa J, Ives AR (2009) Environmental variation in ecological communities and inferences from single-species data. Ecology 90(5):1268–1278

    Article  PubMed  Google Scholar 

  • Akçakaya H, Halley J, Inchausti P (2003) Population-level mechanisms for reddened spectra in ecological time series. J Anim Ecol 72(4):698–702

    Article  Google Scholar 

  • Amarasekare P, Savage V (2012) A framework for elucidating the temperature dependence of fitness. Am Nat 179(2):178–191

    Article  PubMed  Google Scholar 

  • Andrewartha HG, Birch LC (1954) The distribution and abundance of animals. University of Chicago Press, Chicago

    Google Scholar 

  • Ariño A, Pimm S (1995) On the nature of population extremes. Evol Ecol 9:429–443

    Article  Google Scholar 

  • Austin M (2007) Species distribution models and ecological theory: a critical assessment and some possible new approaches. Ecol Model 200(1–2):1–19

    Article  Google Scholar 

  • Bak P, Tang C, Wiesenfeld K (1987) Self-organized criticality: an explanation of the 1/f noise. Phys Rev Lett 59(4):381–384

    Article  PubMed  Google Scholar 

  • Box G, Jenkins G, Reinsel G (2011) Time series analysis: forecasting and control. Wiley, New York

    Google Scholar 

  • Brännström A, Sumpter DJT (2005) The role of competition and clustering in population dynamics. Proc R Soc B Biol Sci 272(1576):2065–2072

    Article  Google Scholar 

  • Caswell H, Cohen JE (1995) Red, White and Blue: environmental Variance spectra and coexistence in metapopulations. J Theor Biol 176:301–316

    Article  Google Scholar 

  • Cohen J (1995) Unexpected dominance of high frequencies in chaotic nonlinear population models. Nature 378(7):610–612

    Article  CAS  PubMed  Google Scholar 

  • Cohen J, Newman C, Cohen A, Petchey O L, Gonzalez A (1999) Spectral mimicry: a method of synthesizing matching time series with different Fourier spectra. Circuits Syst Signal Process 18(3):431–442

    Article  Google Scholar 

  • Cuddington K, Yodzis P (1999) Black noise and population persistence. Proc R Soc B Biol Sci 266:969–973

    Article  Google Scholar 

  • Cyr H (1997) Does inter-annual variability in population density increase with time? Oikos 79(3):549–558

    Article  Google Scholar 

  • Davidson J, Andrewartha H (1948) The influence of rainfall, evaporation and atmospheric temperature on fluctuations in the size of a natural population of Thrips imaginis (Thysanoptera). J Anim Ecol 17(2):200–222

    Article  Google Scholar 

  • de Valpine P, Hastings A (2002) Fitting population models incorporating process noise and observation error. Ecol Monogr 72(1): 57

    Article  Google Scholar 

  • Dennis B, Costantino R (1988) Analysis of steady-state populations with the gamma abundance model: application to Tribolium. Ecology 69(4):1200–1213

    Article  Google Scholar 

  • Dennis B, Otten MRM (2000) Joint effects of density dependence and rainfall on abundance of San Joaquin kit fox. J Wildl Manag 64(2):388–400

    Article  Google Scholar 

  • Dennis B, Ponciano JM, Lele SR, Taper ML, Staples DF (2006) Estimating density dependence, process noise, and observation error. Ecol Monogr 76(3):323–341

    Article  Google Scholar 

  • Engen S, Be Sæther, Armitage K, Blumstein DT, Clutton-Brock T, Dobson F, Festa-Bianchet M, Oli MK, Ozgul A (2013) Estimating the effect of temporally autocorrelated environments on the demography of density independent age structured populations. Methods Ecol Evol 4:573–584

    Article  Google Scholar 

  • Farrior C, Tilman D, Dybzinski R, Reich P, Levin S, Pacala S (2013) Resource limitation in a competitive context determines complex plant responses to experimental resource additions. Ecology 94(11):2505–2517

    Article  PubMed  Google Scholar 

  • Ferguson JM, Ponciano JM (2014) Predicting the process of extinction in experimental microcosms and accounting for interspecific interactions in single-species time series. Ecol Lett 17:251– 259

    Article  PubMed  PubMed Central  Google Scholar 

  • Ferguson JM, Ponciano JM (2015) Evidence and implications of higher-order scaling in the environmental variation of animal population growth. Proc Natl Acad Sci 112(9):2782–2787

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Fowler MS, Ruokolainen L (2013) Confounding environmental colour and distribution shape leads to underestimation of population extinction risk. PLoS One 8(2):e55855

  • Garcia-Carreras B, Reuman D (2011) An empirical link between the spectral colour of climate and the spectral colour of field populations in the context of climate change. J Anim Ecol 80:1042–1048

    Article  PubMed  Google Scholar 

  • Geritz SA, Kisdi E (2004) On the mechanistic underpinning of discrete-time population models with complex dynamics. J Theor Biol 228(2):261–269

    Article  PubMed  Google Scholar 

  • Gonzalez A, Holt RD (2002) The inflationary effects of environmental fluctuations in source–sink systems. Proc Natl Acad Sci 99(23):14:872–14:877

    Article  CAS  Google Scholar 

  • Goodman D (1987) The demography of chance extinction. In: Soule ME (ed) Viable populations for conservation. Cambridge University Press, Cambridge, pp 11–34

    Chapter  Google Scholar 

  • Halley JM (1996) Ecology, evolution and-noise. Trends Ecol Evol 11(1):33–37

    Article  CAS  PubMed  Google Scholar 

  • Halley JM, Inchausti P (2004) The increasing importance of 1/f-noises as models of ecological variability. Fluct Noise Lett 4(2):R1– R26

    Article  Google Scholar 

  • Heino M, Sabadell M (2003) Influence of coloured noise on the extinction risk in structured population models. Biodivers Conserv 110(3):325

    Google Scholar 

  • Hilfinger A, Paulsson J (2011) Separating intrinsic from extrinsic fluctuations in dynamic biological systems. Proc Natl Acad Sci USA 12(29):167–172

    Google Scholar 

  • Holt R (2009) Bringing the Hutchinsonian niche into the 21st century: ecological and evolutionary perspectives. Proc Natl Acad Sci USA 106(2):19:659–19:665

    Article  CAS  Google Scholar 

  • Holt RD, Barfield M, Gonzalez A (2003) Impacts of environmental variability in open populations and communities: “inflation” in sink environments. Theor Popul Biol 64(3):315–330

    Article  PubMed  Google Scholar 

  • Hooker HD (1917) Liebig’s law of the minimum in relation to general biological problems. Science 46(1183):197

    Article  PubMed  Google Scholar 

  • Hosking J (1981) Fractional differencing. Biometrika 68(1):165– 176

    Article  Google Scholar 

  • Huey R, Stevenson R (1979) Integrating thermal physiology and ecology of ectotherms: a discussion of approaches. Am Zool 366:357–366

    Article  Google Scholar 

  • Inchausti P, Halley J (2001) Investigating long-term ecological variability using the global population dynamics database. Science 293(5530):655–657

    Article  CAS  PubMed  Google Scholar 

  • Inchausti P, Halley J (2002) The long-term temporal variability and spectral colour of animal populations. Evol Ecol Res 4:1033–1048

    Google Scholar 

  • Johnson J (1925) The Schottky effect in low frequency circuits. Phys Rev 541(1918):71–85

    Article  Google Scholar 

  • Jonzén N, Lundberg P (2002) The irreducible uncertainty of the demography–environment interaction in ecology. Proc R Soc B Biol Sci 269:221–225

    Article  Google Scholar 

  • Jonzén N, Pople T, Knape J, Sköld M (2010) Stochastic demography and population dynamics in the red kangaroo Macropus rufus. J Anim Ecol 79(1):109–116

    Article  PubMed  Google Scholar 

  • Kaitala V, Ylikarjula J, Ranta E, Lundberg P (1997) Population dynamics and the colour of environmental noise. Proc R Soc B Biol Sci 264(1384):943–948

    Article  CAS  Google Scholar 

  • Kamenev A, Meerson B, Shklovskii B (2008) How colored environmental noise affects population extinction. Phys Rev Lett 101(268103)

  • Kearney M, Phillips BL, Tracy CR, Christian KA, Betts G, Porter WP (2008) Modelling species distributions without using species distributions: the cane toad in Australia under current and future climates. Ecography 31:423–434

    Article  Google Scholar 

  • Knape J, de Valpine P (2010) Effects of weather and climate on the dynamics of animal population time series. Proc R Soc B Biol Sci 278(1708):985

    Article  Google Scholar 

  • Knape J, de Valpine P (2011) Are patterns of density dependence in the Global Population Dynamics Database driven by uncertainty about population abundance? Ecol Lett 15(1):17–23

    Article  PubMed  Google Scholar 

  • Laakso J, Kaitala V, Ranta E (2001) How does environmental variation translate into biological processes? Oikos 92(1):119–122

    Article  Google Scholar 

  • Laakso J, Kaitala V, Ranta E (2003a) Non-linear biological responses to disturbance: consequences on population dynamics. Ecol Model 162(3):247–258

    Article  Google Scholar 

  • Laakso J, Löytynoja K, Kaitala V (2003b) Environmental noise and population dynamics of the ciliated protozoa Tetrahymena thermophila in aquatic microcosms. Oikos 102:663–671

    Article  Google Scholar 

  • Laakso J, Kaitala V, Ranta E (2004) Non-linear biological responses to environmental noise affect population extinction risk. Oikos 104:142–148

    Article  Google Scholar 

  • Lande R (1993) Risks of population extinction from demographic and environmental stochasticity and random catastrophes. Am Nat 142(6):911–927

    Article  Google Scholar 

  • Lewontin RC, Cohen D (1969) On population growth in a randomly varying environment. Proc Natl Acad Sci USA 62(4):1056–1060

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Lindén A, Knape J (2009) Estimating environmental effects on population dynamics: consequences of observation error. Oikos 118:675–680

    Article  Google Scholar 

  • Lindén A, Fowler M, Jonzén N (2013) Mischaracterising density dependence biases estimated effects of coloured covariates on population dynamics. Popul Ecol 55:183–192

    Article  Google Scholar 

  • Lögdberg F, Wennergren U (2012) Spectral color, synchrony, and extinction risk. Theor Ecol

  • Miramontes O, Rohani P (1998) Intrinsically generated coloured noise in laboratory insect populations. Proc R Soc B Biol Sci 265:785–792

    Article  Google Scholar 

  • Mode CJ, Jacobson ME (1987) A study of the impact of environmental stochasticity on extinction probabilities by Monte Carlo integration. Math Biosci 83(1):105–125

    Article  Google Scholar 

  • Montroll E, Shlesinger M (1982) On 1/f noise and other distributions with long tails. Proc Natl Acad Sci USA 79(10):3380–3383

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Morales JM (1999) Viability in a pink environment: why “white noise” models can be dangerous. Ecol Lett 2(4):228–232

    Article  Google Scholar 

  • Morris R (1959) Single-factor analysis in population dynamics. Ecology 40(4):580–588

    Article  Google Scholar 

  • Murdoch WW, Kendall BE, Nisbet RM, Briggs CJ, McCauley E, Bolser R (2002) Single-species models for many-species food webs. Nature 417(6888):541–543

    Article  CAS  PubMed  Google Scholar 

  • NERC (2010) The Global Population Dynamics Database Version 2

  • Petchey O (2000) Environmental colour affects aspects of single-species population dynamics. Proc R Soc B Biol Sci 267(1445):747–754

    Article  CAS  Google Scholar 

  • Petchey O, Gonzalez A, Wilson H (1997) Effects on population persistence: the interaction between environmental noise colour, intraspecific competition and space. Proc R Soc B 264:1841–1847

    Article  PubMed Central  Google Scholar 

  • Pike N, Tully T, Haccou P, Ferrière R (2004) The effect of autocorrelation in environmental variability on the persistence of populations: an experimental test. Proc R Soc Lond Ser B Biol Sci 271(1553):2143

    Article  Google Scholar 

  • Pimm SL, Redfearn A (1988) The variability of population densities. Nature 334:613–614

    Article  Google Scholar 

  • Pinheiro J, Bates D, DebRoy S, Sarkar D, R Development Core Team (2011) nlme: linear and nonlinear mixed effects models

  • R Development Core Team (2012) R: a language and environment for statistical computing

  • Ranta E, Lundberg P, Kaitala V, Laakso J (2000) Visibility of the environmental noise modulating population dynamics. Proc Biol Sci / The Royal Society 267(1455):1851–1856

    Article  CAS  Google Scholar 

  • Ratikainen II, Ja Gill, Gunnarsson TG, Sutherland WJ, Kokko H (2008) When density dependence is not instantaneous: theoretical developments and management implications. Ecol Lett 11(2):184–198

    PubMed  Google Scholar 

  • Ricker W (1954) Stock and Recruitment. J Fish Res Board Can 11(5):559–623

    Article  Google Scholar 

  • Ripa J, Lundberg P (1996) Noise colour and the risk of population extinctions. Proc R Soc B Biol Sc 263(1377):1751–1753

  • Rotenberg M (1987) Effect of certain stochastic parameters on extinction and harvested populations. J Theor Biol 124:455–471

    Article  Google Scholar 

  • Roughgarden J (1975) A simple model for population dynamics in stochastic environments. Am Nat 109 (970):713–736

    Article  Google Scholar 

  • Roy M, Holt RD, Barfield M (2005) Temporal autocorrelation can enhance the persistence and abundance of metapopulations comprised of coupled sinks. Am Nat 166(2):246–261

    Article  PubMed  Google Scholar 

  • Royama T (1977) Population persistence and density dependence. Ecol Monogr 47(1):1–35

    Article  Google Scholar 

  • Royama T (1981) Fundamental concepts and methodology for the analysis of animal population dynamics, with particular reference to univoltine species. Ecol Monogr 51(4):473–493

    Article  Google Scholar 

  • Royama T (1992) Analytical population dynamics. Chapman & Hall, London

    Book  Google Scholar 

  • Ruokolainen L, McCann K (2013) Environmental weakening of trophic interactions drives stability in stochastic food webs. J Theor Biol 339:36–46

    Article  PubMed  Google Scholar 

  • Ruokolainen L, Fowler MS, Ranta E (2007) Extinctions in competitive communities forced by coloured environmental variation. Oikos 116:439–448

    Article  Google Scholar 

  • Ruokolainen L, Lindén A, Kaitala V, Fowler MS (2009) Ecological and evolutionary dynamics under coloured environmental variation. Trends Ecol Evol 24(10):555–563

    Article  PubMed  Google Scholar 

  • Savage VM, Gillooly JF, Brown JH, West GB, Charnov EL (2004) Effects of body size and temperature on population growth. Am Nat 163(3):429–441

    Article  PubMed  Google Scholar 

  • Schwager M, Johst K, Jeltsch F (2006) Does red noise increase or decrease extinction risk? Single extreme events versus series of unfavorable conditions. Am Nat 167(6):879–888

    Article  PubMed  Google Scholar 

  • Shaffer M (1987) Minimum viable populations: coping with uncertainty. In: Soule M (ed) Viable populations for conservation, chap Minimum vi. Cambridge University Press, Cambridge, pp 69–86

    Chapter  Google Scholar 

  • Shumway R, Stoffer D (2006) Time series analysis and its applications. Springer, New York

    Google Scholar 

  • Sibly RM, Barker D, Denham MC, Hone J, Pagel M (2005) On the regulation of populations of mammals, birds, fish, and insects. Science 309(5734):607–610

    Article  CAS  PubMed  Google Scholar 

  • Sibly RM, Barker D, Hone J, Pagel M (2007) On the stability of populations of mammals, birds, fish and insects. Ecol Lett 10:970–976

    Article  PubMed  Google Scholar 

  • Staples DF, Taper ML, Dennis B (2004) Estimating population trend and process variation for PVA in the presence of sampling error. Ecology 85(4):923–929

    Article  Google Scholar 

  • Sugihara G (1995) From out of the blue. Nature 378(7):559–60

    Article  CAS  Google Scholar 

  • Swanson B (1998) Autocorrelated rates of change in animal populations and their relationship to precipitation. Conserv Biol 12(4):801–808

    Article  Google Scholar 

  • Taper M, Gogan P (2002) The northern Yellowstone elk: density dependence and climatic conditions. J Wildl Manag 66(1):106–122

    Article  Google Scholar 

  • Taper ML, Staples DF, Shepard BB (2008) Model structure adequacy analysis: selecting models on the basis of their ability to answer scientific questions. Synthese 163(3):357–370

    Article  Google Scholar 

  • Tuljapurkar S (1982) Population dynamics in variable environments. II. Correlated environments, sensitivity analysis and dynamics. Theor Popul Biol 140:114–140

    Article  Google Scholar 

  • Tuljapurkar S, Haridas CV (2006) Temporal autocorrelation and stochastic population growth. Ecol Lett 9(3):327–337

    Article  PubMed  Google Scholar 

  • van de Pol M, Vindenes Y, Sæther BE, Engen S, Ens BJ, Oosterbeek K, Tinbergen JM (2011) Poor environmental tracking can make extinction risk insensitive to the colour of environmental noise. Proc R Soc B Biol Sci 278(1725):3713–3722

    Article  Google Scholar 

  • Vasseur DA, Yodzis P (2004) The color of environmental noise. Ecology 85(4):1146–1152

    Article  Google Scholar 

  • Veenstra JQ (2012) Persistence and anti-persistence: theory and software. PhD thesis, Western University

  • Wagenmakers EJ, Farrell S, Ratcliff R (2004) Estimation and interpretation of 1/f noise in human cognition. Psychon Bull Rev 11(4):579–615

    Article  PubMed  PubMed Central  Google Scholar 

  • Wichmann M, Johst K, Moloney K, Wissel C, Jeltsch F (2003) Extinction risk in periodically fluctuating environments. Ecol Model 167:221–231

    Article  Google Scholar 

  • Wichmann MC, Johst K, Schwager M, Blasius B, Jeltsch F (2005) Extinction risk, coloured noise and the scaling of variance. Theor Popul Biol 68(1):29–40

    Article  PubMed  Google Scholar 

  • Ziebarth NL, Abbott KC, Ives AR (2010) Weak population regulation in ecological time series. Ecol Lett 13(1):21–31

    Article  PubMed  Google Scholar 

Download references

Acknowledgments

We thank the associate editor and two anonymous reviewers for their thoughtful reviews which greatly improved the quality of this manuscript. We thank Robert Holt, John Hopkins, and Craig Osenberg for reviewing an earlier version of this manuscript and the class, Zoology 6927, Quantitative Methods in Ecology, at the University of Florida for making this project possible. We would like to the lab of Colette St. Mary for constructive comments on this manuscript. JMP was supported by the National Institute of General Medical Sciences of the National Institutes of Health under Grant R01-GM103604. JMF gratefully acknowledges the support by the National Science Foundation under Grant No. 0801544 in the Quantitative Spatial Ecology, Evolution and Environment Program at the University of Florida and the Global Population Dynamics database: http://www3.imperial.ac.uk/cpb/databases/gpdd.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Jake M. Ferguson.

Appendices

Appendix A: Curating the GPDD

We chose a high quality subset of the over 5000 GPDD datasets using the following criteria: there must have been at least 15 observations in the time series, the qualitative GPDD reliability rating must have been 4 or 5 (out of a maximum rating of 5), and the data must not have been constant over the first 3 years. In addition, we only allowed datasets where sampling units indicated nonharvest indices, as harvests may not reflect true population abundances. All data were transformed by adding 1 to all observations, in order to remove any 0’s. We tested the effects of this data transformation by analyzing a subset of the 445 datasets where this transformation was not applied, but fewer datasets (166) were available for this analysis. The MainID’s of the GPDD datasets used in our analysis were 1 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 56 57 58 59 60 61 62 63 64 65 66 67 1088 1089 1090 1091 1092 1093 1094 1095 1096 1097 1098 1099 1100 1101 1102 1103 1104 1105 1106 1107 1108 1109 1110 1111 1112 1113 1114 1115 1116 1172 1173 1174 1175 1176 1177 1178 1179 1180 1181 1182 1183 1184 1185 1186 1187 1188 1189 1211 1213 1248 1250 1254 1261 1268 1272 1275 1276 1347 2770 2771 2772 2773 2774 2775 2776 2777 2778 2779 2780 2781 2787 2789 2791 2798 2800 2802 2805 2810 2815 2829 2833 2840 2842 2844 2852 2857 2867 2869 2879 2887 2891 2901 2903 2905 2915 2917 2922 2926 2935 2937 2939 2958 2960 2962 2974 2976 2986 2991 3001 3003 3017 3021 3028 3030 3037 3043 3045 3051 3056 3059 3068 3070 3072 3084 3092 3108 3110 3112 3118 3123 3127 3133 3135 3139 3144 3157 3159 3161 3170 3177 3182 3190 3214 3216 3218 3223 3225 3233 3237 3249 3251 3253 3260 3263 3265 3268 3275 3283 3295 3297 3306 3312 3316 3323 3330 3356 3358 3360 3372 3378 3382 3393 3395 3406 3426 3428 3430 3437 3442 3445 3455 3466 3468 3470 3477 3480 3482 3484 3489 3496 3508 3521 3533 3535 3539 3546 3557 3559 3567 3575 3577 3583 3585 3592 3623 3625 3627 3639 3641 3650 3664 3666 3673 3676 3678 3680 3683 3688 3693 3706 3708 3716 3718 3723 3726 3739 3741 3757 3774 3776 3784 3787 3795 3797 3799 3811 3814 3825 3827 3829 3838 3840 3849 3853 3864 3866 3876 3882 5019 5020 6057 6104 6105 6106 6107 6108 6109 6110 6111 6112 6113 6114 6115 6116 6117 6118 6119 6120 6121 6122 6123 6124 6125 6126 6127 6128 6129 6130 6131 6132 6133 6134 6135 6136 6137 6138 6139 6140 6141 6142 6143 6592 6728 6729 6730 6731 6732 6733 6734 6735 6736 6737 6738 6739 6740 6741 6742 6743 6744 6745 6788 6789 6790 6791 6792 6793 6794 6795 6796 6797 6798 6799 6800 6801 6802 6803 6804 6805 6806 6807 6808 6809 6810 6811 6812 6813 6814 6815 6816 6817 6818 6819 6820 6821 6822 6823 6824 6825 6826 6827 6828 6829 6830 6831 6832 6833 6834 6835 6836 6837 6838 6839 6840 6841 6842 6843 6844 6845 6846 6847 6848 6849 6850 6851 6852 6853 6854 6855 6856 6857 6858 6859 6860 6861 6862 7065 9191 9192 9193 9245 9246 9247 9248 9249 9279 9280 9302 9338 9339 9340 9341 9342 9343 9344 9345 9346 9347 9348 9365 9391 9393 9425 9517 9518 9519 9536 9542 9684 9833 9834 9911

Appendix B: Parameter estimation and model selection

Parameter estimation

Parameter estimation was performed using both maximum likelihood (ML) and restricted maximum likelihood (ReML) methods. We used one-step predictions to fit population dynamics models with PEA using abundance data from the GPDD. AICc was used to perform model selection from among the candidate models using the model’s ML estimates. However, our reported values of \(\hat {\phi }\) used ReML parameter estimates from the best AICc models. ReML estimation has been shown to perform better than traditional ML when estimating variance components, but is not valid for model selection using information criterion (Staples et al. 2004). ML and ReML estimations were performed using the generalized least squares (gls) function from the NLME package in the R software environment (Pinheiro et al. 2011; R Development Core Team 2012). We assumed a multivariate normal distribution for the observed per capita growth rate, \(\ln \left (\frac {N(t)}{N(t-1)}\right )\), where the mean vector was given by the predicted per capita growth rate for the corresponding form of density dependence. The covariance matrix was given by σ 2 R, where the correlation matrix, R, was given by the correlation structure for the appropriate ARMA(1,1) model used. For k observations of the per capita growth rate, the ARMA(1,1) correlation structure with AR(1) parameter ϕ and MA(1) parameter θ is given by

$$ \mathbf{R} = \frac{(1 + \theta\phi)(\theta + \phi)}{1 + 2\theta\phi + \theta^{2}}\left[ \begin{array}{llll} 1 & \phi^{0} & {\ldots} & \phi^{k-2}\\ \phi^{0} & 1 & {\ldots} & \phi^{k-3}\\ {\vdots} & {\vdots} & {\ddots} & \vdots\\ \phi^{k-2} & \phi^{k-3} & {\ldots} & 1 \end{array} \right], $$
(B1)

where the AR(1) model corresponds to a special case where θ = 0. For example, the log-likelihood for the Gompertz lag 1 density dependence model was then given by,

$$\begin{array}{@{}rcl@{}} \ln(L) &=& -\frac{k \ln(2\pi)}{2} - \frac{k \ln\left| \sigma^{2}\mathbf{R} \right|}{2} - \\ &&\frac{1}{2\sigma^{2}}\left( \ln \left( \frac{N(t)}{N(t-1)}\right) - a + b \ln(N(t-1)) \right)^{T}\\ &&\times \mathbf{R}^{-1} \left( \ln \left( \frac{N(t)}{N(t-1)}\right) - a + b \ln\left( N(t-1)\right) \right).\\ \end{array} $$
(B2)

Appendix C: Covariate analysis

We tested the explanatory power of a suite of covariates, available in the GPDD, on \(\hat {\phi }\). We explored whether environmental factors, species information, and data quality metrics could explain variation in \(\hat {\phi }\). Specifically, we tested whether time series length, latitude, and longitude of the data collection location, GPDD reliability index, and sampling frequency were significant predictors of \(\hat {\phi }\) at the α = 0.10 level. We ran an additional regression looking at species taxa as predictors of \(\hat {\phi }\) at the α = 0.10 level. We also tested whether species status as specialists or generalists could explain variation in \(\hat {\phi }\). We used definitions from Murdoch et al. (2002) to determine specialist/generalist status in 48 time of the GPDD series. We predicted that generalists, with potentially more trophic interactions would have lower levels of PEA than specialists who may have fewer, but stronger, trophic interactions consistent with our results from the SLLM model in the main text.

Results

Our exploratory regression analysis on \(\hat {\phi }\) was used to determine whether covariates could explain results from our analysis. We used the arctanh transformation on \(\hat {\phi }\), defined by \(Z=\text {arctan} (\hat {\phi })\), in order to satisfy the assumption of normally distributed residuals. We performed a test on all sampling units reported in the GPDD. We found that only the Breeding females category was a statistically significantly predictor (t test, p value = 0.0293), though when using the Bonferonni correction for multiple tests this is no longer a significant predictor. The two species in this category were Tringa nebularia and Melospiza melodia. We also examined whether sample size and GPDD reliability index influenced estimates. We found that \(\hat {\phi }\) tended to decrease with increases in sample size and the reliability index, though these effects were not statistically significant.

Status as specialist (\(\hat {\phi }=0.11\)) or generalist (\(\hat {\phi }=0.03\)) was not found to be a significant explanatory value of \(\hat {\phi }\) at the α = 0.1 level (p value = 0.44). However, the results were consistent with our predictions with the specialist having slightly higher mean levels of autocorrelation, consistent with predictions from the SLLM.

Appendix D: Alternative error models

Past work has formulated autocorrelation models several different ways. Here, we discuss the differences in two alternative model formulations of temporally autocorrelated processes and contrast their properties to methods used by past studies examining the TPA.

A potentially important distinction in autocorrelated time series models is between long- and short-memory processes. Short-memory processes have autocorrelation functions that decay exponentially to 0 as k (e.g., exp(−ϕ k)), while long-memory processes have autocorrelation functions that converge to 0 according to slow-decaying power-law functions as k (e.g., k β) (Shumway and Stoffer 2006). Thus, though both processes may have the same degree of autocorrelation and go to 0, short-memory processes go to 0 faster than long-memory models. Short-memory models depend only on recent realizations of the process while long-memory models exhibit autocorrelation over many past realizations of the process. The relevance of this distinction is that incorporating PEA in long- and short-term memory models with the same degree of autocorrelation can lead to different predictions about species persistence in otherwise identical models (Cuddington and Yodzis 1999; Fowler and Ruokolainen 2013). These findings suggest that distinguishing between long- and short-memory PEA may have practical implications for population modeling.

While it is often convenient to think of population abundances as functions of time, previous discussions of long-memory processes in the ecological literature have often used the frequency domain representation of time series (Halley 1996; Cuddington and Yodzis 1999; Vasseur and Yodzis 2004). This approach decomposes a time dependent signal into an infinite sum of sine waves with different frequencies through the Fourier transform. The transformation calculates the amplitude for each sine wave, giving the relative contribution of that frequency to the original series. This frequency representation can then be used as a convenient diagnostic tool when determining the appropriateness of a particular time dependent model (Box et al. 2011).

The power spectral density function, denoted as S(f), gives the distribution of the frequencies f present in a time series. The most commonly used long-memory process is the inverse power law model (1/f model) (Johnson 1925). For this model, the spectral density function is given by

$$ S(f)=\text{constant} /f^{\beta}, $$
(D1)

where β controls the degree of autocorrelation. The 1/f model has also been proposed several times as a general model of environmental variation for population dynamics due to its apparent ubiquity in natural phenomena (Montroll and Shlesinger 1982; Bak et al. 1987; Halley 1996). The mathematical representation of the autocorrelation in a time series model and the function S(f) are transforms of each other and thus are mathematically equivalent.

Long-memory models are typically investigated within the framework of frequency domain methods while short memory processes are usually studied using traditional time domain tools, which tend to be more convenient for model estimation. We used the fractionally differenced model (FDM) to model time series with long-memory. In the FDM, the current state of the autocorrelated process is given as the sum of all the past contributions to the process and an independent random shock (Hosking 1981). Autocorrelation is controlled by a single coefficient, d, which determines the relative contributions of all past states. The mathematical representation of the model is

$$ E(t) = \sum\limits_{k=1}^{\infty} \frac{-{\Gamma}(k-d)}{\Gamma(k+1){\Gamma}(-d)}E(t-k)+ W(t), $$
(D2)

where E(t) is the state of a process at time t, Γ(x) is a gamma function, d is the fractional differencing parameter where −0.5<d<0.5, and W(t) is a normal distribution with mean equal to zero and variance of σ 2. In this model, the state at the current time step depends on all previous time lags through an infinite series representation, leading to the long-memory property. The differencing degree d controls the degree of autocorrelation and corresponds approximately to β/2 in the 1/f model (Hosking 1981). Thus, the FDM and the 1/f model are approximately equivalent for −1<β<1.

Because previous methods have often used frequency domain approaches, we determined whether estimates of the total population autocorrelation (TPA) made to the GPDD dataset using an AR(1) model would be different from previous studies that were based on the estimated spectral exponent, a long-memory model used in several past studies of TPA (Pimm and Redfearn 1988; Inchausti and Halley 2001). We were also interested in whether the FDM model would be more consistent with the long-memory process that is spectral exponent than the AR(1) model. We first estimated the spectral exponent by fitting a linear regression to the Fourier transform of the log-transformed abundances in our GPDD dataset consistent with previous methods. The transform was calculated with the fft function in the R software environment (R Development Core Team 2012). The frequency amplitude of the transformed time series, S(f), was then fit to a model of the form, ln(S(f))=aβln(f, where f is the observed sampling frequency of the time series signal, β is the spectral exponent, and a is an intercept term. We also estimated the FDM and AR(1) coefficients in Eqs. D2 and 3 using the arfima package in R (Veenstra 2012) from the log-transformed abundance time series in order to compare them to \(\hat {\beta }\).

We evaluated the consistency of the AR(1) model and the FDM with \(\hat {\beta }\) by regressing each models autocorrelation estimates (ϕ in Eq. 3 and d in Eq. D2) against \(\hat {\beta }\) using standard major axis linear regression due to the higher variance in estimates of β. For the AR(1) model, we fit, \(\hat {\beta }/2 = a_{\phi } + b_{\phi } \hat {\phi } + W\), and for the FDM model we fit \(\hat {\beta }/2 = a_{d} + b_{d} \hat {d} + W\), where W was a random variable assumed to be iid normally distributed. Consistency with the 1/f β model was determined by whether the confidence intervals of the intercept parameters (a) contained 0 and the intervals of the slope parameters (b) contained 1. This comparative method was used rather than directly performing model selection on AR(1) and FDM models of PEA due to the inability for information theoretic methods to reliably distinguish between long- and short-memory models (Wagenmakers et al. 2004). We also determined if the properties of the data sets used here were consistent with previous studies on the GPDD by comparing our estimates of the TPA in our GPDD datasets using the 1/f model to previous estimates of TPA in the GPDD (e.g., Inchausti and Halley 2001).

Results

We found levels of the spectral exponent with our dataset, \(\hat {\beta }=0.94\) that were similar to previous estimates of \(\hat {\beta }=1.02\) (Inchausti and Halley 2002), indicating that the data set used here is comparable to previous work on the GPDD. We also compared autocorrelation estimates from both the AR(1) and FDM error processes to \(\hat {\beta }\) in order to determine model consistency with previous estimates from the GPDD. In the comparison of the AR(1) and to 1/f noise, we did not include estimates that indicated nonstationary time series, which led to 387 data sets in the analysis. In the comparison of the FDM model and 1/f model, this same criterion led to 244 data sets being included in the analysis. For the AR(1) model, we found an intercept value of \(\hat {a}_{\phi }=-0.0.072\), 95 % CI =(−0.101,−0.045), and slope of \(\hat {b}_{\phi }=1.18\), 95 % CI = (1.109, 1.26). For the FDM model, we found an intercept value of \(\hat {a}_{d}=0.075\), 95 % CI = (0.069, 0.081), and slope of \(\hat {b}_{d}=0.086\), 95 % CI =(0.78,0.95). These results indicate that there may be bias in both models with regards to comparisons with the spectral exponent. However, the AR(1) model does make estimates that are broadly consistent with the spectral exponent, an interesting result. Despite the important differences in the properties of the spectral density of the AR(1) and the spectral exponent, our results suggest that parameter estimation using the AR(1) model is comparable to estimates that would be obtained with a long-memory process population environmental autocorrelation (PEA) model.

Appendix E: PEA in the carrying capacity

We performed a simulation analysis in order to determine if our analysis would provide biased estimates when the PEA occurs in the carrying capacity rather the maximum per capita rate of increase. The carrying capacity in the lag 1 Ricker model, given by K = a/b in Eq. 2, was assumed to be an AR(1) random variable. We set a = ln(1.5) and E[K]=100 for all simulations. We tested 15 different autocorrelation levels evenly spaced in the interval [−0.95, 0.95] over three different levels of variability in K, where variation was measured used the coefficient of variation (standard deviation of K divided by the mean). The CV values we tested were 0.5, 1, and 2. For values of the CV higher than 2, we found that some values of our simulated K became negative and were therefore unrealistic. For each combination of autocorrelation and CV value, we estimated the autocorrelation from a time series with length of 105 following the methods of the main text that assume that the variation occurs in the intrinsic growth rate.

We found that estimation error was negligible at the low CV levels of 0.5 and 1 (Fig. 6). At CV =2, there was estimation bias at very low negative levels of the autocorrelation but the bias was low for positive levels of autocorrelation. Overall, we found that the influence of this kind of model specification to be low, suggesting that our results are not biased by the way that the environmental variation enters into the population dynamics. These results are not too surprising as variation in the carrying capacity is often difficult to distinguish from variation in the maximum per capita rate of increase with differences emerging primarily at very low or very high abundances (Ferguson, unpublished results).

Fig. 6
figure 6

The impact on estimated PEA when the PEA arises in the population carrying capacity but the error model is in the intrinsic rate of reproduction. The degree of estimation error increases as a function of the coefficient of variation (CV) in the carrying capacity and is larger for negative values of the PEA

Appendix F: Spectral mimicry

In the main text, we used the known distributional form of stationary AR time series to ensure that the simulated time series had identical properties except for the degree of autocorrelation. Although this allowed us to control all the statistical moments, alternative methods, such as spectral mimicry, exist that allow the generation of time series using the same realization of the generated data with only the degree of autocorrelation changing. We used this method to check if our approach properly accounted for changes in statistical properties with autocorrelation.

Spectral mimicry (Cohen et al. 1999) can vary the autocorrelation in the same realization of a stochastic process, thereby controlling for the effects of not scaling of process variance with autocorrelation or other unintended mismatches that may arise between simulated series. The method works by first generating a series from a stochastic model such as a standard normal distribution. In order to turn this iid stochastic process into an autocorrelated one, a new reference series with the desired degree of autocorrelation is generated. The original process is then reordered such that it’s sequence of order statistics matches the reference series. This reordered version has the same realized values as the original process, however, the autocorrelation is equal to the reference series. Because the new series has the same values as the original series, just in a different order, the statistical properties of the ensemble must be equivalent.

We reproduced Figs. 2 and 3 using datasets generated with spectral mimicry in order to confirm that our method of generating autocorrelated data did not lead to any differences between time series other than the degree of autocorrelation. The reproduced figures (Figs. 7 and 8) show no differences with the figures in the main text (Figs. 2 and 3).

Fig. 7
figure 7

The effect of two different nonlinear interactions on an autocorrelated time series, C(t), when data is generated using the spectral mimicry method. The dashed black line is the autocorrelation in the untransformed environmental covariate, while the colored lines are the population environmental autocorrelation (PEA) values in E(t) for different transformations. Panel a corresponds to power transforms of the form E(t)=−C(t)η. Panel b corresponds to the logistic function, \(E(t) = \frac {e^{\zeta C(t)}}{1 + e^{\zeta C(t)}}\)

Fig. 8
figure 8

Two applications of the SLLM to a number, n, of autocorrelated time-series. In both panels the x-axis denotes the autocorrelation in the environmental covariate while y-axis is the population environmental autocorrelation (PEA) that affects the per capita growth rate once the SLLM has been applied. In panel a, all n covariates have the same autocorrelation. In panel b, n−1 of the covariates have an autocorrelation of 0.5, while the remaining covariate has the autocorrelation value given on the x-axis. In both panels, the dashed line denotes the covariate’s autocorrelation, for comparison. The overall effect of the SLLM depends on the number of environmental covariates, n, that limit the population. In general, the SLLM tends to reduce the magnitude of the PEA relative to what is expected from a single limiting factor. However, as shown in the right panel, the effect can increase the PEA if a single limiting environmental factor has a high enough autocorrelation

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Ferguson, J.M., Carvalho, F., Murillo-García, O. et al. An updated perspective on the role of environmental autocorrelation in animal populations. Theor Ecol 9, 129–148 (2016). https://doi.org/10.1007/s12080-015-0276-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s12080-015-0276-6

Keywords

Navigation