UC Santa Barbara UC Santa Barbara Previously Published Works

Title Dynamical effects of plant quality and parasitism on population cycles of larch budmoth

Permalink https://escholarship.org/uc/item/3q39z2cn

Journal , 84(5)

Authors Turchin, Peter Wood, Simon N Ellner, Stephen Paul et al.

Publication Date 2003

DOI 10.1890/0012-9658(2003)084[1207:DEOPQA]2.0.CO;2

Peer reviewed

eScholarship.org Powered by the California Digital Library University of California Ecology, 84(5), 2003, pp. 1207±1214 ᭧ 2003 by the Ecological Society of America

DYNAMICAL EFFECTS OF PLANT QUALITY AND PARASITISM ON POPULATION CYCLES OF LARCH BUDMOTH

PETER TURCHIN,1 SIMON N. WOOD,2 STEPHEN P. E LLNER,3 BRUCE E. KENDALL,4 WILLIAM W. M URDOCH,5 ANDREAS FISCHLIN,6 JEÃ ROME CASAS,7 EDWARD MCCAULEY,8 AND CHERYL J. BRIGGS9 1Department of Ecology and Evolutionary , University of Connecticut, Storrs, Connecticut 06269 USA 2Mathematical Institute, North Haugh, St. Andrews Fife, KY16 9SS, UK 3Department of Ecology and Evolutionary Biology, Corson Hall, Cornell University, Ithaca, New York 14853 USA 4School for Environmental Science and Management, University of California, Santa Barbara, California 93106 USA 5Department of Ecology, Evolution, and Marine Biology, University of California, Santa Barbara, California 93106 USA 6Institute of Terrestrial Ecology, Swiss Federal Institute of Technology Zurich ETHZ, CH-8952 Schlieren/Zurich, Switzerland 7University of Tours, IRBI-CNRS UPRESA 6035, F-37200 Tours, France 8Ecology Division, Department of Biological Sciences, University of Calgary, Calgary, Canada T2N 1N4 9Department of Integrative Biology, University of California, Berkeley, California 94720 USA

Abstract. Population cycles have been remarkably resistant to explanation, in part because crucial experiments are rarely possible on appropriate spatial and temporal scales. Here we show how new approaches to nonlinear time-series analysis can distinguish between competing hypotheses for population cycles of larch budmoth in the Swiss Alps: delayed effects of budmoth density on food quality, and budmoth±parasitoid interactions. We re- examined data on budmoth density, plant quality, and parasitism rates. Our results suggest that the effect of plant quality on budmoth density is weak. By contrast, a simple model of budmoth±parasitoid interaction accounts for 90% of the variance in budmoth population growth rates. Thus, contrary to previous studies, we ®nd that parasitoid±budmoth interaction appears to be the dominant factor driving the budmoth cycle. Key words: forest insects; larch budmoth; nonlinear time-series analysis; parasitism; plant qual- ity; population cycles; Switzerland; trajectory matching; Zeiraphera diniana.

INTRODUCTION hand, Kendall et al. (1999) showed, for a laboratory population, how a combination of time-series analysis of forest insects have been a and mechanistic modeling could serve as an equally focus of numerous controversies in ecology. One area of contention is whether outbreak cycles are driven by valid approach for distinguishing between alternative density dependent processes or by exogenous forces hypotheses to explain population cycles. In this paper (Royama 1981, Turchin 1990, Hunter and Price 1998). we show that this approach can be applied successfully Within the density-dependence camp, arguments rage to populations in the ®eld, to explain the cause of cycles between proponents of intrinsic factors, e.g., maternal in a forest insect pest, the larch budmoth (LBM). effects (Ginzburg and Taneyhill 1994) vs. trophic hy- Larch budmoth (Zeiraphera diniana) populations in potheses (Berryman 1995). the Swiss Alps are among the most remarkable re- Because dynamics occur on large spatial and tem- corded cycles, oscillating with a period of 8±9 yr and poral scales, in many cases it is not feasible, or at least ϳ100 000-fold change between peak and trough den- it is very dif®cult and expensive, to test hypotheses sity (Fig. 1). Ecological theory offers several potential about dynamic mechanisms experimentally. This is explanations for such regular population cycles in for- probably why there are so few experimental studies est insects: parasitoid±host interaction (Hassell 1978), that successfully tested mechanistic hypotheses of pop- delayed effects of plant quality (Fischlin and Balten- ulation cycles, and only one, as far as we know, in a sweiler 1979, Edelstein-Keshet and Rausher 1989), forest insect (snowshoe hares, Krebs et al. 1995; Red pathogen±host interaction (Anderson and May 1980), Grouse, Hudson et al. 1998; voles, KorpimaÈki and and maternal effects (Ginzburg and Taneyhill 1994). A Norrdahl 1998; southern pine beetles, Turchin et al. large empirical research effort has been devoted to 1999). If the only methodologically valid approach for studying the LBM cycle and trying to identify the caus- elucidating the causes of population cycles is manip- al mechanisms (see review in Baltensweiler and Fis- ulative experiments in the ®eld, then we should expect chlin 1988). As a result of this effort, the last two very slow progress in this area of ecology. On the other hypotheses can be rejected in this system. There is no empirical support for maternal effects, and although the granulosis virus infection was observed to cause Manuscript received 28 March 2002; revised 11 September 2002; accepted 17 September 2002; ®nal version received 2002. substantial mortality during the ®rst two intensively Corresponding Editor: F. R. Adler. studied outbreaks (50% in the peak year of 1955 and 1207 1208 PETER TURCHIN ET AL. Ecology, Vol. 84, No. 5

FIG. 1. Population ¯uctuations of larch bud- moth density in the Upper Engadine Valley (for explanation of ``Engadine'' vs. ``Sils'' data, see Methods: Sources of data).

25% in the 1964 peak), subsequent outbreaks collapsed Data were collected at multiple sites throughout the without being accompanied by any viral mortality (Bal- valley separately (Auer 1977; these data are tabulated tensweiler and Fischlin 1988). in Fischlin 1982), but because all sites oscillated in Plant quality (e.g., raw ®ber and protein content) and close synchrony, we can average them into one time parasitism have the necessary delayed effects to induce series that we call ``Engadine.'' The density of budmoth cycles. It takes two or more years for foliage quality larvae (third instar) is expressed in number per unit of to recover after heavy defoliation, which accompanies 1 kg of branches with foliage. For the period of 1952± a budmoth population peak. Furthermore, ®eld and lab- 1976 (with one year, 1968, missing) we have data on oratory bioassays show that poor food quality has a the percentage of larvae parasitized, also averaged over strong effect on budmoth survival and reproduction multiple sites (Delucchi 1982) omitting several data (Benz 1974, Omlin 1977, Fischlin 1982). Thus, the points that were interpolated by the author. Although current explanation of budmoth cycles is based on their numerous parasitoids are associated with larch bud- interaction with foliage quality (Baltensweiler and Fis- moth, parasitism is dominated by the ichneumonid Phy- chlin 1988, Berryman 1999). todietus griseanae and three eulophid species, whose Measured parasitism rates vary between lows of 1± ¯uctuations are correlated (R2 ϭ 0.94 between log Phy- 5% and highs of 80±90% (Delucchi 1982), and max- todietus abundance and log eulophid abundance, P Ͻ imum parasitism rates are achieved ϳ2 yr after bud- 0.001). After 1977, sampling in the Upper Engadine moth peaks. Previous studies concluded that mortality valley continued on a reduced scale. At one site, Sils, by parasitoid wasps does not cause the cycle, but mere- data were collected in an uninterrupted sequence from ly tracks the budmoth population (Delucchi 1982, Bal- 1951 to 1992: we refer to these as the ``Sils'' data tensweiler and Fischlin 1988). However, as our analysis (Baltensweiler 1993; A. Fischlin, unpublished data). will show, this rejection of the parasitoid hypothesis During the period of 1961±1992, needle lengths of trees was inappropriate. at Sils were also measured. Needle length is a good The goal of our paper is to empirically distinguish index of plant quality because it is well correlated with between the two rival hypotheses for which there is raw ®ber and protein content of larch needles (Omlin empirical evidence in the larch budmoth (LBM) sys- 1977, Fischlin 1982). Furthermore, bioassay data of tem: plant quality vs. parasitism. To do this, we for- Benz (1974) indicated that needle length has a strong mulated dynamic models of the LBM system that em- effect on larval survival and pupal mass (and pupal body the rival hypotheses. We then use these mecha- mass is closely connected to adult fecundity). Turchin nism-based models as statistical models for the purpose (2003) calculated that the needle length of foliage with of testing hypotheses about which mechanism is best which LBM larvae were fed in the Benz (1974) bio- supported by the available data. Conceptually, we are assays explained 86% variance in a measure of LBM simply using the standard approach to statistical hy- ®tness (the product of larval survival and adult fecun- pothesis testing, but the highly nonlinear of the dity). models complicates the implementation of the ap- proach, as described below in Methods: Models. Models

METHODS To decrease the chance that an inappropriate mod- eling choice would bias our results against the plant Sources of data quality hypothesis, we modeled the effect of plant qual- Systematic population census of larch budmoth in ity on budmoth dynamics with two alternative func- the Upper Engadine Valley (Switzerland) started in tional forms. The ®rst (the ``nonlinear'' version) is a 1949 and with minor modi®cations continued until modi®ed Ricker model in which the discrete rate of 1977 (Auer 1977, Baltensweiler and Fischlin 1988) budmoth population increase is a hyperbolic (saturat- (lesser quality data are also available for 1945±1948). ing) function of plant quality (for equations, see Table May 2003 HOST QUALITY, PARASITISM IN POPULATION CYCLES 1209

TABLE 1. Dynamic models and model equations for larch budmoth populations in Switzerland.

Model and parameters Equation Plant quality I (the ``nonlinear'' version) Qt Nt : budmoth density Ntϩ10ϭ␭ Nttexp[ϪgN ] ␦ϩQt

Nt Qt : plant quality index Qtϩ1 ϭ (1 Ϫ␣)1Ϫϩ␣Qt ΂΃␥ϩNt

Plant quality II (alternative Nt equation, the ``linear'' version)

Nt : budmoth density Ntϩ1 ϭ Ntttexp[u ϩ vL Ϫ gN ] Lt : needle length Parasitoid±host aPt Nt : budmoth density Ntϩ10ϭ␭ Nttexp ϪgN Ϫ []1 ϩ ahNttϩ awP

aPt Pt : parasitoid density Ptϩ1 ϭ␾Nt 1 Ϫ exp Ϫ Ά·[]1 ϩ ahNttϩ awP

Tritrophic I (full model) QaPtt Nt : budmoth density Ntϩ10ϭ␭ Nttexp ϪgN Ϫ ␦ϩQttt[]1 ϩ ahN ϩ awP

aPt Pt : parasitoid density Ptϩ1 ϭ␾Nt 1 Ϫ exp Ϫ Ά·[]1 ϩ ahNttϩ awP

Nt Qt : plant quality index Qtϩ1 ϭ (1 Ϫ␣)1Ϫϩ␣Qt ΂΃␥ϩNt

Tritrophic II (simpli®ed ``linear'' version) N : budmoth density t aPt Pt : parasitoid density Ntϩ1 ϭ Nttexp u ϩ vL Ϫ []1 ϩ awPt Lt : needle length

Notes: All state variables and parameters are bounded below by 0. In addition, Qt, ␣, ␾, ␥, h, and w are bounded above by 1. The proportion of larvae parasitized was computed as St ϭ [1 Ϫ exp{ϪaPt /(1 ϩ ahPt ϩ awNt)}]. We assumed that the ®eld data underestimate (some instances of parasitism are missed because ®eld-collected larvae must be reared in the lab) so the expected observed parasitism rate was assumed to be ␥St where ␥ Յ 1 is a parameter to be ®tted. The measure of plant quality in our data is needle length, and we assumed a linear relationship between needle length and actual plant quality, Qt ϭ b ϩ cLt, where b and c are parameters to be ®tted.

1). The parameter g within the exponent quanti®es di- ``linear'' version of the plant-quality model with the rect density dependence. The second (the ``linear'' ver- simpli®ed form of parasitoid functional response (all sion) assumes that the realized per capita rate of bud- equations in Table 1). moth population change (rt ϭ log(Nt/NtϪ1) is a linear function of the plant quality index and LBM density. General philosophy of analysis In the plant quality equation, quality next year is de- Our approach is based on the standard logic of sta- creased by current-year herbivory in a hyperbolic fash- tistical hypothesis testing. However, because our hy- ion. Once herbivore damage becomes negligible (as a potheses are rich in mechanism, the mathematical mod- result of collapse of the budmoth population) quality els embodying these hypotheses are substantially more increases gradually to the maximum, and the speed of nonlinear and detailed than is usual in statistical hy- recovery is regulated by the parameter ␣. pothesis testing. As noted in the Introduction, we are The host±parasitoid model is based on the standard focusing on two ``monocausal'' hypotheses (food qual- Nicholson-Bailey framework, but with self-regulation ity and parasitism). In addition, it is possible that a in the host and a general form of functional response composite explanation is correct (both food and par- (Beddington 1975) that combines effects of prey sat- asitism jointly explain the LBM cycle). Once we have uration and parasitoid mutual interference. translated all hypotheses into models, we ask whether The tritrophic model simply combines the nonlinear the observed data could plausibly have been generated version of the plant quality model and the parasitoid± by either of the simpler models or whether we require host model. In addition, we investigated a simpli®ed the more complicated model to explain them. The pro- version of the tritrophic model, which combined the cess involves ®tting each model to the observed data 1210 PETER TURCHIN ET AL. Ecology, Vol. 84, No. 5

(estimating its parameters), so that model comparisons measure model ®t) can show a good deal of structure, can be made with each model performing as well as including local minima, at a variety of scales. Much of its basic structure will allow (of course, both ®tting and this variability is of no statistical signi®cance, in the comparison must be performed using the same measure sense that it is of a scale below that of the sampling of ®t). This is precisely the approach that underlies variability of the ®tting objective itself. Nonetheless, almost all statistical hypothesis testing. such ®ne-scale structure can make model ®tting dif®- The fact that standard statistical methods follow the cult and tends to spoil the performance of the large- framework that we use here can be seen by considering sample approximations such as Likelihood Ratio tests the theory underlying a familiar one-way ANOVA. In that would usually allow for statistical inferences about this case two models are compared, a null model that the ®tted models. We deal with this issue by using the data are all observations of random variables having bootstrapping methods, both as a means of escaping the same mean, and an alternative model that they are from small-scale local minima during ®tting (see Wood observations of random variables whose mean depends 2001a) and for statistical comparison of models. on the level of some factor. Both models are ®t to the Formally, then, we treat each model comparison as data by least squares, and their ®t compared by the testing a null hypothesis that a simple model is correct difference in their residual sum of squares. The P value against an alternative that a more complicated model of the ANOVA is obtained by considering the distri- is needed. The sequence of steps is as follows: bution of difference in sum of squares that would be 1) Both models are ®tted to data, and their difference expected if the null model were true: we have to aban- in ®t, ⌬, is noted. don the simpler model only if the improvement in mod- 2) We create replicate simulated data sets under the el ®t of the alternative is larger than could plausibly null hypothesis that the simpler model is true by boot- have occurred if the null model is true. strap resampling the residuals from the simpler model The models that we want to compare are highly non- and adding this resample to the ®tted values from the linear and this causes some dif®culty in formulation simpler model. and ®tting. In particular, we must deal with two types 3) Both models are ®tted to each bootstrap-simulated of stochasticity: process error and measurement error. data set and the difference in their ®t ⌬* is noted for Process error is variability due to factors that are not each replicate. explicitly modeled (``noise''). It affects the true values 4) The ⌬ value is compared to the distribution of ⌬* of population density. Measurement error is variability values generated by simulationÐif it lies well within due to imprecisely measuring population density. Si- the distribution then the null hypothesis that the simple multaneous treatment of both types of variability with model is correct can be accepted, otherwise it will be highly nonlinear models is a relatively new develop- rejected in favor of the alternative model. A P value ment, requiring computationally intensive methods that can be obtained by assessing the proportion of ⌬* val- have not yet received much testing (Ellner et al. 2002). ues which are more favorable to the alternative model We therefore chose to work with two complementary than the observed ⌬. measures of ®t, each addressing only one source of error. The ®rst measure of model ®t (one-step-ahead One-step-ahead ®tting ®tting) considers the average ability of the model to The response variable was de®ned as the realized per predict the population change over a year, given mea- capita rate of population change, de®ned as rt ϭ ln(Nt) surements of the system state at the beginning of the Ϫ ln(NtϪ1) (Turchin 1990, Berryman 1991, Royama year: this amounts to neglecting measurement error. 1992). Each model used (see Table 1) could be recast The second measure of model ®t (global ®tting, also to predict this quantity given measurements of some known as ``trajectory matching'') measures the ability or all of NtϪ1, LtϪ1, and PtϪ1(LBM density, needle length of the model to predict the entire series of data when [an index of food quality], and parasitism rate, respec- iterated forward deterministically from starting con- tively), depending on the model in question. Model ditions, which are also estimated by ®tting: this parameters were estimated by minimizing the mean amounts to neglecting process error, but yields accurate square difference between model predictions of rt and estimates if the period of oscillations is nonetheless observed rt using the nonlinear regression routines in constant or nearly so, as in LBM (Kendall et al. 1999, STATISTICA (Statsoft 1999). Wood 2001b). If conclusions based on both measures As explained in Methods: General philosophy of concur then we can have some con®dence that neither analysis, bootstrap resampling of the model residuals of their key approximations is critical (Hilborn and was performed by resampling with replacement from Mangel 1997). the model residuals. We selected the subset of years A key practical issue (especially for the trajectory- (1962±1978) for which we had both needle length data matching approaches) is that model ®tting and infer- and parasitism data. We used Sils budmoth data as the ence are not straightforward for highly nonlinear, pa- response variable in the regression analysis (thus, this rameter-rich models when using relatively small data procedure biases our results in favor of the plant quality sets. In particular, the model ®tting objectives (which hypothesis). Furthermore, we used the linear model rt May 2003 HOST QUALITY, PARASITISM IN POPULATION CYCLES 1211

TABLE 2. Results of analyses of time-series data. ϭ u ϩ vLtϪ1, because our analysis showed that it ®tted the data better than the alternative model in Table 1 a) Nonlinear regression (given the one-step-ahead ®tting objective). We also Model Dependent variable R2 ®tted the data with the parasitism model rt ϭ r0 Ϫ aPtϪ1/

(1 ϩ awPtϪ1), and the tritrophic model rt ϭ u ϩ vLtϪ1 Plant quality larch budmoth (rt) 0.312 Ϫ aP /(1 ϩ awP ) (this is Tritrophic Model II in Plant quality plant quality (Qt) 0.615 tϪ1 tϪ1 Parasitoid±host larch budmoth (r ) 0.880 Table 1). Note that the test we constructed is conser- t Parasitoid±host parasitoid (Pt) 0.712 vative with respect to ®nding for the parasitism hy- b) Results from trajectory matching pothesis. Model R2 P value Global ®tting Full tritrophic 0.839 ´´´ No feedback from plant 0.828 0.025 Global ®tting was performed using the methods de- quality to host scribed in Wood (2001b). Each model was ®tted to all Plant quality only 0.443 Ͻ0.003 2 available data predicted by that model, using a least Notes: R is the coef®cient of determination, and rt ϭ log squares objective. In this case the model is not making Nt ϩ 1/Nt. Other variables and equations are the same as in predictions of the next year's population change given Table 1. observations of state variables at the start of the year and the parameters, but is predicting each of the entire series of data given the model parameters and some dynamics. The linear model ®ts the data better than the starting conditions (which like the parameters, must be nonlinear alternative, but still explains only 31% of the estimated). To stabilize variances, the LBM density was variance in the realized per capita rate of change of raised to the power 0.1 for ®tting, and a linear trans- budmoth, rt ϭ ln(Nt) Ϫ ln(NtϪ1) (Table 2a; the nonlinear formation was applied to needle length in order that model explains 15% of the variance). Interestingly, the needle length residuals were comparable to trans- last documented budmoth peak (in 1989, not shown in formed LBM and parasitism rate residuals. Parasitism our graphs) was curtailed due to several years of un- rates were ®tted on the usual arcsine square-root trans- favorable weather in winter and spring (Baltensweiler formed scale. Bootstrapping of the residuals was per- 1993). As a result, budmoth did not achieve the usual formed by resampling with replacement on a series by peak density and there was no widespread defoliation. series basis. Local minima were avoided during ®tting The plant quality index did not collapse, but budmoth by bootstrap restarting (Wood 2001a). density did, suggesting that poor plant quality (as mea- Again, the computer intensive methods described in sured by needle length) may not be necessary for driv- the section Methods: General philosophy of analysis ing the budmoth population to a low level. were used to test reduced models against the tritrophic By contrast, parasitism rates explain almost 90% of model, using difference in residual sum of squares be- variance in the per capita rate of change of budmoth tween the alternative models as ⌬. First, all four time numbers (Table 2a). In fact, a simple model rt ϭ r Ϫ series (Engadine, Sils, needle length, and parasitism aPtϪ1/(1 ϩ awPtϪ1) with only three ®tted parameters rate) were ®tted with the trajectory-matching method explains the bulk of the effect of parasitism on budmoth (1) by the tritrophic model and (2) by a reduced model rate of population change (R2 ϭ 86.5%). Using a dif- in which food quality has no in¯uence on the host but ferent metric, log Nt instead of rt, and the same model, hosts have an impact on food quality (i.e., ␦ϭ0 in the we ®nd that 94.1% of the variance in log Nt is explained tritrophic model of Table 1). The effect of plant quality by parasitism rates (see Fig. 2). We are not aware of was modeled using the ®rst alternative in Table 1, be- any other ®eld data sets in in which cause we found that it ®tted data better than the linear such a simple model can explain a comparable amount model. Again, this procedure biases results in favor of of variation. Interestingly, the effects of budmoth den- the plant quality model. Secondly, the three series other sity on plant quality and on parasitism rates are almost than the parasitism rate were ®tted (1) by the full mod- equally strong, with Ͼ70% of variance explained in el, and (2) by the plant quality model of Table 1. (The each case (Table 2a). plant quality model cannot be used to predict parasitism The analysis of joint effects of quality and parasitism rate: when we remove from the model the effects of in the reduced data set showed that the plant quality parasitism on LBM, we also remove the effects of LBM model explained 48.0% of the variance in budmoth rate on the parasites because both are functions of the same of population change (the higher proportion of variance parasitism rate). explained than in the full data set is primarily due to the reduced data set not including the last oscillation). RESULTS The parasitism and tritrophic models explained 65.6% and 66.1% of variance, respectively. Exploratory regression analysis The trajectory matching (global ®tting) analysis The regression analysis indicated that the plant qual- yielded very similar results. The full tritrophic model ity index has a surprisingly weak effect on budmoth provided an excellent ®t to the data (Fig. 3), marginally 1212 PETER TURCHIN ET AL. Ecology, Vol. 84, No. 5

DISCUSSION Summary of LBM results Our re-examination of data on budmoth density, plant quality, and parasitism rates suggests that the ef- fect of plant quality on budmoth density is weak. By contrast, a simple model of budmoth±parasitoid inter- action accounts for 90% of the variance in budmoth population growth rates. Thus, contrary to previous studies, we ®nd that parasitoid±budmoth interaction ap- pears to be the dominant factor driving the budmoth cycle. The fact that parasitism does not peak until bud- moth is already in decline was previously taken as ev- FIG. 2. Comparing predictions of the ®tted parasitoid model to data. Predictions are obtained according to the fol- idence against the parasitism hypothesis (Delucchi à lowing formula: Nt ϭ NtϪ1 exp[r Ϫ aStϪ1NtϪ1/(1 ϩ awStϪ1NtϪ1)] 1982). However, our model shows that this phase lag where Nt and St are observed budmoth density and proportion is consistent with parasitism driving the cycles (Fig. parasitized, and r, a, and w are ®tted parameters. 3). Our results, however, should not be interpreted as better then the model omitting the effect of plant quality suggesting that plant quality plays no role in larch bud- on budmoth (Table 2b). By contrast, the food quality moth cycles. Although the regression approach could model ®tted poorly (Table 2b). not detect a statistically signi®cant effect of plant qual- To summarize, exploratory one-step-ahead regres- ity on larch budmoth population rate of change, the sion analysis indicated that parasitoids have a strong trajectory matching results indicate that the tritrophic effect on budmoth dynamics, but it did not detect any model ®ts data slightly (and statistically signi®cantly, effect of plant quality. Global ®ts suggested similar P ϭ 0.025) better than the parasitism-only model. results albeit hinting at the possibility of a slight effect The fact that the regression (one-step-ahead ®tting) of plant quality. These results are consistent with the and trajectory matching (global ®tting) approaches are hypothesis that budmoth population cycles are driven not in perfect agreement about the effect of plant qual- primarily (or solely) by the moth's interactions with ity re¯ects the different model ®tting criteria, which parasitoids, with plant quality ¯uctuating in response have different strengths and weaknesses. The regres- to changes of budmoth numbers. sion analysis deals well with process noise by concen- trating on short time scales, but consequently has low Formal hypothesis testing power to detect weaker interactions whose effects on Using one-step-ahead ®tting the parametric bootstrap the dynamics only become apparent over longer time test, described in Methods, indicated that the tritrophic model ®tted data signi®cantly better than plant quality model (P ϭ 0.029). By contrast, the parasitism model ®tted the data as well as the tritrophic model (P ϭ 0.64). In other words, with this measure of ®t, the anal- ysis yields no evidence for an effect of plant quality on the rate of larch budmoth population growth. This result is particularly striking because our statistical test was designed to be biased in favor of plant quality hypothesis (see Methods: Models). Using global ®tting and testing the full tritrophic model against the null hypothesis of a reduced ``par- asitoid only'' model (in which the feedback from plant quality to budmoth is eliminated) yielded an estimated P value of 0.025, indicating that the small effect of plant quality on host dynamics seen in the one-step- ahead regression analysis is nonetheless real and plays a role in shaping budmoth cycles. Testing the full model against the plant-quality-only model, yielded an esti- mates of P Ͻ 0.003 with this measure of ®t. Thus, there is strong evidence that plant quality alone cannot ac- FIG. 3. Results of trajectory matching. Upper panel: En- gadine data set (hollow symbols indicate budmoth density; count for the observed host cycles, which lends ®rm ®lled symbols indicate proportion parasitized). Lower panel: support to both the exploratory results and the formal Sils data set (hollow symbols indicate budmoth density; ®lled testing with the one-step-ahead approach. symbols indicate needle length). May 2003 HOST QUALITY, PARASITISM IN POPULATION CYCLES 1213 scales. It is also sensitive to measurement errors, which associated with rejecting the food quality hypothesis, are usually nontrivial in any ecological ®eld study. In see Table 2) should be regarded as a decisive result. contrast, trajectory matching deals well with measure- The same force of evidence emerging from a dynamical ment error noise, and is sensitive to long time-scale model should be equally decisive. In fact, we would patterns in the data, but makes an unrealistic assump- argue that it should count for more; our result is based tion that process noise is absent. The observation that on biologically plausible and empirically grounded dy- the two approaches agree on the role of parasitism in namical models, while the assumptions of ANOVA are larch budmoth dynamics greatly strengthens this result. rarely subject to scrutiny. Another feature of our approach that strengthens the main conclusion is that we did everything we could to ACKNOWLEDGMENTS avoid biasing the analysis against the food quality hy- This paper is a result of a collaborative effort under the pothesis. In fact, we consistently made choices that auspices of the Complex Population Dynamics working group at the National Center for Ecological Synthesis and Analysis, could be regarded as biasing the analysis in favor of Santa Barbara, CA, USA. NCEAS is funded by NSF (grant the food-quality hypothesis, and nonetheless it failed number DEB-94-21535), the University of California-Santa to perform better than the parasitism hypothesis. Barbara, and the State of California. We thank Werner Bal- tensweiler for help with data sources and valuable criticism. General implications Alan Berryman provided valuable critique.

Models based on very different ecological mecha- LITERATURE CITED nisms can produce dynamics that are qualitatively sim- Anderson, R. M., and R. M. May. 1980. Infectious diseases ilar (Hunter and Dwyer 1998). For example, pathogen, and population cycles of forest insects. Science 210:658± plant quality, and parasitism models can all generate 661. second-order oscillations (that is, cycles characterized Auer, C. 1977. Dynamik von LaÈrchenwicklerpopulationen by delayed density dependence) with periods of 7±10 laÈngs des Alpenbogens. Mitteilungen der EidgenoÈssischen Anstalt fuÈr forstliches Versuchswesen 53:71±105. yr typical of forest insect populations. Traditional ap- Baltensweiler, W. 1993. Why the larch bud-moth cycle col- proaches using qualitative comparisons between model lapsed in the subalpine larch±cembran pine forests in the predictions and data are unable to distinguish between year 1990 for the ®rst time since 1850. Oecologia 94:62± dynamically plausible alternative hypotheses. Further- 66. more, since experimental manipulations at appropriate Baltensweiler, W., and A. Fischlin. 1988. The larch budmoth in the Alps. Pages 331±351 in A. A. Berryman, editor. spatiotemporal scales are often impossible or prohib- Dynamics of forest insect populations: patterns, causes, itively expensive, it would seem that we are unable to implications. Plenum Press, New York, New York, USA. make further progress in elucidating the causes of for- Beddington, J. R. 1975. Mutual interference between para- est insect cycles. sites or predators and its effect on searching ef®ciency. Journal of Animal Ecology 44:331±340. In a previous paper (Kendall et al. 1999), we argued Benz, G. 1974. Negative Ruckkopelung durch Raum- und that a way out of this apparent impasse is to develop Nahrungskonkurrenz sowie zyklische Veranderung der approaches that combine mechanistic modeling with Nahrungsgrundlage als Regelsprinzip in der Popultions- rigorous time-series statistics. Kendall et al. (1999) dynamik des Grauen Larchenwicklers, Zeiraphera diniana showed that in a laboratory setting such a synthetic (Guenee) (Lepidoptera, Tortricidae). Zeitschrift fuer an- gewandte Entomologie 76:196±228. approach can distinguish among rival mechanisms that Berryman, A. A. 1991. Population theory: an essential in- predict qualitatively similar dynamics. In this paper, gredient in pest prediction, management and policy making. we show that the same approach can be successfully American Entomologist 37:138±142. used in a ®eld situation. The clear message from our Berryman, A. A. 1995. Population cycles: a critique of ma- ternal effects and allometric hypotheses. Journal of Animal analysis is that the dynamic data best support the hy- Ecology 64:290±293. pothesis that LBM cycles are driven by parasitism, but Berryman, A. A. 1999. Principles of population dynamics modi®ed slightly by an interaction with food quality. and their application. Stanley Thornes, Cheltenham, UK. The reason we can obtain such a strong result is because Delucchi, V. 1982. Parasitoids and hyperparasitoids of Zeir- we utilize the data more fully (performing detailed aphera diniana (Lepidoptera Tortricidae) and their role in population control in outbreak areas. Entomophaga 27:77± quantitative comparison of models with data) and be- 92. cause we employ not only data on LBM dynamics, but Edelstein-Keshet, L., and M. D. Rausher. 1989. The effects also on the dynamics of potential interactors (food of inducible plant defenses on herbivore populations. I. quality and parasitoids). Mobile herbivores in continuous time. American Naturalist 133:787±810. Although technically quite involved, our approach Ellner, S. P., Y. Seifu, and R. H. Smith. 2002. Fitting pop- has the same logical structure as standard hypothesis ulation dynamics models to time-series data by gradient tests such as ANOVA: two models (e.g., effect present matching. Ecology 83:2256±2270. vs. effect absent) are compared based on their ability Fischlin, A. 1982. Analyse eines wald-insekten-systems: der to ®t the data, and the simpler is rejected in favor of subalpine LaÈrchen-arvenwald und der graue LaÈrchenwick- ler Zeiraphera diniana Gn (Lepidoptera, Tortricidae). Dis- the more complex if the improvement in goodness of sertation Number 6977. ETH, Zurich, Switzerland. ®t is large enough. Few ecologists would dispute that Fischlin, A., and W. Baltensweiler. 1979. Systems analysis an ANOVA table with P Ͻ 0.003 (this is the P value of the larch bud moth system. Part 1: the larch±larch bud 1214 PETER TURCHIN ET AL. Ecology, Vol. 84, No. 5

moth relationship. Mitteilungen der Schweizerischen En- 1995. Impact of food and predation on the snowshoe hare tomologischen Gesellschaft 52:273±289. cycle. Science 269:1112±1115. Ginzburg, L. R., and D. E. Taneyhill. 1994. Population cycles Omlin, F. X. 1977. Zur populationsdynamischen Wirkung der of forest Lepidoptera: a maternal effect hypothesis. Journal durch Raupenfrass und Dungung veranderten Nahrungs- of Animal Ecology 63:79±92. basis auf den Grauen Larchenwickler Zeiraphera diniana Hassell, M. P. 1978. The dynamics of arthropod predator± Gn. (Lep: Tortricidae). Dissertation Nr. 6064. ETH Zurich, prey systems. Princeton University Press, Princeton, New Switzerland. Jersey, USA. Royama, T. 1981. Fundamental concepts and methodology Hilborn, R., and M. Mangel. 1997. The ecological detective. for the analysis of animal population dynamics, with par- Princeton University Press, Princeton, New Jersey, USA. ticular reference to univoltine species. Ecological Mono- Hudson, P. J., A. P. Dobson, and D. Newborn. 1998. Pre- graphs 51:473±493. vention of population cycles by parasite removal. Science Royama, T. 1992. Analytical population dynamics. Chapman 282:2256±2258. and Hall, London, UK. Hunter, A. F., and G. Dwyer. 1998. Outbreaks and interactive StatSoft. 1999. STATISTICA for Windows. StatSoft, Tulsa, factors: insect population explosions synthesized and dis- Oklahoma, USA. sected. Integrative Biology 1:166±177. Turchin, P. 1990. Rarity of density dependence or population Hunter, M. D., and P. W. Price. 1998. Cycles in insect pop- regulation with lags? Nature 344:660±663. ulations: delayed density dependence or exogenous driving variables? Ecological Entomology 23:216±222. Turchin, P. 2003. Complex population dynamics: a theoret- Kendall, B. E., C. J. Briggs, W. W. Murdoch, P. Turchin, S. ical/empirical synthesis. Princeton University Press, P. Ellner, E. McCauley, R. M. Nisbet, and S. N. Wood. Princeton, New Jersey, USA. 1999. Why do populations cycle? A synthesis of statistical Turchin, P., A. D. Taylor, and J. D. Reeve. 1999. Dynamical and mechanistic modeling approaches. Ecology 80:1789± role of predators in population cycles of a forest insect: an 1805. experimental test. Science 285:1068±1071. KorpimaÈki, E., and K. Norrdahl. 1998. Experimental reduc- Wood, S. N. 2001a. Minimising model ®tting objectives that tion of predators reverses the crash phase of small-rodent contain spurious local minima by bootstrap restarting. Bio- cycles. Ecology 79:2448±2455. metrics 57:240±244. Krebs, C. J., S. Boutin, R. Boonstra, A. R. E. Sinclair, J. N. Wood, S. N. 2001b. Partially speci®ed ecological models. M. Smith, M. R. T. Dale, K. Martin, and R. Turkington. Ecological Monographs 71:1±25.