w OPTIMALITY, CONSTRAINTS, AND KIERARCKIES IN THE ANALYSIS OF FORAGING STRATEGIES '^' ;-■ V t ' BY JEFFREY ROBERT LUCAS A DISSERTATION PRESENTED TO THE GRADUATE COUNCIL OF THE UNIVERSITY OF FLORIDA IK PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF DOCTOR OF PHILOSOPHY UNIVERSITY OF FLORIDA 1985 ^- ACKNOWLEDGMENTS '• This dissertation is the end result of the input and help from many X people. Dr. James Dufty derived the models I used in the first chapter, and Dr. Karl Taylor verified the math. Linda Griffin helped a great deal with Chapter III, and weighed lots of fruit flies and fed lots of antlions. Steven Frank, Alan Grafen, and Dr. Eric Charnov were excellent sounding boards for some of my ideas on the theory of foraging behavior; Steve and Alan were also instrumental in teaching me all I know about statistical models. Dr. Franl: Mature provided a fellowship that gave me time to finish the work on Chapter I and also provided easy access to Sea Horse Key. Dr. Mike LaBarbera was a god-send in showing me what biomechanics is, and also in explaining the ins and outs of Stoke 's Law. Thanks to my committee, Drs. Jane Brockmann, Carmine Lanciani, Brian McKab, Frank Nordlie, and Howard T. Odum, for their aid and comments on a slightly non-traditional dissertation. The following people read and commented on one or more chapters: Dr. Robert Jaeger, Dr. Michael LaBarbera, Dr. A. Richard Palmer, Lynda Peterson, Dr. Nancy Stamp, Dr. Lionel Stange, Dr. Nat Wheelwright, Steven Frank, Alan Grafen, Dr. Lincoln Brower, Dr. Mark Denny, and several anonym.ous reviewers. Most of my ideas about systems are derived from the work of Dr. Odum, to wnom I am indebted. Special thanks go to my wife, Lynda Peterson, for many different things. Finally, this work reflects an il pose incredible effort on the part of my chairperson, Jane Brockmann, who worked nearly as hard on this dissertation as I did, and whose ideas are integrated in every chapter. Ill TABLE OF CONTENTS PAGE ACKNOWLEDGEMENTS ii ABSTRACT vi INTRODUCTION 1 CHAPTER I - THE BIOPHYSICS OF PIT CONSTRUCTION 5 Introduction 5 General Methods 6 Pit-Construction Behavior 7 Pit Morphology and Prey Behavior 9 Physical Components of Pit Construction 9 Behavioral Components of Pit Construction 23 General Discussion 35 CHAPTER II - MODELS OF PARTIAL PREY CONSUMPTION 38 Introduction 38 Proximate Models 39 Optimal Foraging Models 42 Capture Probability and Ambush Predation 60 Discussion 62 CHAPTER III - PARTIAL PREY CONSUMPTION BY ANTLION LARVAE. 68 Introduction 68 Digestion Rate Limitation (DRL) Model 69 Deterministic Optimality Model 72 Stochastic Optimality Model 95 General Discussion 107 CHAPTER IV - THE ROLE OF FORAGING TIME CONSTRAINTS AND VARIABLE PREY ENCOUNTER IN OPTIMAL DIET CHOICE 110 Introduction 110 The Cost Model 112 Discussion lAO Summary 145 IV CHAPTER V - OPTIMALITY, HIERARCHIES AND FORAGING 148 Introduction 148 Optimality 149 Hierarchy 153 Maximum Power and Foraging Hierarchies 156 Non-Hierarchical Foraging Models 159 Hierarchy and Optimality Models 163 CONCLUSIONS 170 LITERATURE CITED 173 BIOGRAPHICAL SKETCH 182 Abstract of Dissertation Presented to the Graduate Council of the University of Florida in Partial Fulfillment of the Requirements for the Degree of Doctor of Philosophy OPTIMALITY, CONSTRAINTS, AND HIERARCHIES IN THE ANALYSIS OF FORAGING STRATEGIES By Jeffrey Robert Lucas April, 1985 Chairperson: H. Jane Brockmann Major Department: Zoology Three phases of foraging behavior were analyzed: (l) the preparation for prey (specifically trap construction by antlions), (2) diet choice, and (3) consumption of prey. Optimal foraging models were formulated for each phase. Results suggest that behavioral modeling should be constructed in a hierarchical framework. (1) Antlion larvae were shown to line their pitfall traps with fine sand, which significantly increases capture ef ficiencj''. This fine-sand layer is caused by physical properties of sand (angle of repose and sand trajectory), and by three components of pit-construction behavior: regulation of trajectory angle and initial velocity, and pre-sorting of thrown sand. (2) Two general models were derived that predict diet choice when foraging time is unconstrained and when external factors constrain f oraging-bout length. For a two-prey system, the forager should specialize when K >E^h ./(E . h .-E .h. ) , where i,j=high and vi .s^.^ low quality prey (respectively), E=energy, h=handling time. M_. . is the number of i missed while handling j and is shown to correlate with relative cost of eating i- When foraging- bout length is constrained such that one prey is taken per bout, K. . no longer measures the cost of eating i. Here the predator should specialize when P(Y.E.P(Y.)/E. where P(Y . )=probability that i arrives in a foraging bout and P(Y .^ Figure 1.1. Steps of pit construction by antlion larvae. (A) Random movement. (B) Beginning of circular movement; sand thrown to outside of circle. (C) Antlion continues to circle inward in a spiralling path, creating a furrow. (D) Finished pit. common in sandy areas, but is utilized to some degree in any habitat type. Pit Morphology and Prey Behavior If an antlion constructs a pit in sand consisting of several grain sizes (which is the usual case in nature), the completed pit is generally lined with the finest sand available (Fig. 1.2). To test whether this feature functions to increase the efficiency of the pit, I constructed artificial pits by pressing conical molds into sand of different grain sizes. Escape time from these pits was measured for two species of ants, the carpenter ant (Camponotus floridanus) and the smaller fire ant (Solenopsis invicta). Ant escape time increased significantly with decreasing sand grain size for both ant species (Table 1.1, Figs 1.3 and 1.4). Therefore an antlion pit lined with the finest sand available should serve as a more efficient trap than a pit lined with the unsorted spectrum of available sand. Two other variables also significantly affected ant escape time: pit diameter and slope (Table 1.1). Two components may affect the distribution of sand in the pit: (l) the physical properties of sand as they apply to pit morphology, and (2) the behavioral aspects of pit construction. These components are examined separately in the following sections. Physical Components of Pit Construction Pit morphology is affected by two different physical processes. The first determines the "behavior" of sand on the furrow walls during Figure 1.2. Photograph of a completed antlion pit showing the distribution of fine (white) and coarse (black) sand grains. (a) Pit wall lined with white sand. (B) Position of black sand 'ring'. The white line marks the pit edge. 11 12 i»- u lij in 2 UJ o t/5 CAMPONQTUS PtT DIAMETER | • • 35 min O- O 50 mm Ct -A 65 mm F 1. Pit diameter 1 2. Slope 1 3. Sand size *** 3 4. Regression model including all variables 5 5. Error 475 34.35 0.0001 102.60 0.0001 51.04 0.0001 58.06 0.0001 Camponotus f loridanus Independent variables degrees of freedom F value Prob.>F 1. Pit diameter 1 2. Slope 1 3. Sand size *** 3 4. Regression model including all variables 5 5. Error 235 127.88 0.0001 20.72 0.0001 25.61 0.0001 45.08 0.0001 All regression analyses were run on SAS computer program GLM (Barr et al. 1979). Each data point represents a different individual. The following values of the independent variables were used: Solenopsis and Camponotus: pit diameter: 35, 50 and 65 mm; sand ** grain size: 125-250, 250-500, 500-1000 and 1000-2000 um; Solenopsis slope: 30, 35, 40 and 45 ; Camponotus slope: 35 and 40®"; *** Sand size was entered as a class variable and therefore is treated as a non-continuous variable with four levels. -^ 15 construction. This process will directly affect the morphology of the pit and is closely related to the angle of repose (as discussed below). The second process governs the trajectory of thrown particles. Particle trajectory indirectly influences pit morphology in that it will determine what types of sand particles leave the pit after being thrown. Slope: Angle of Repose The antlion pit is lined with fine sand even before it is completed (Fig. 1.5). During construction the larger particles tend to fall to the bottom of the furrow where the animal is digging, leaving the furrow walls lined with finer sand. The differential response of particles of different sizes on the furrow walls suggests that particle size may in itself affect the distribution of sand on the slope. To test this, I constructed artificial pits of different grain sizes. This was done by drawing sand through a hole in the bottom of a tray filled with sorted sand. The slope of the pit walls constructed in this way reflects the angle of repose of the sand. The angle of repose (0') is the maximum slope that sand will attain without collapsing. A significant negative correlation was obtained between sand grain size and slope (r =-0.771, N=34, P<0.01). Thus, for the sand in which the antlion was making a pit, larger particles had a lower 0' and therefore were more likely to fall off a slope than smaller ones. Although a significant correlation was demonstrated, this correlation may reflect differences in particle angularity, roughness, or water content which can covary with particle size. All these factors will affect the angle of internal friction (0, the minimal angle of ?i^-: Figure 1.5. Photograph of an antlion pit in construction showing the distribution of fine (white) and coarse (black) sand grains. (A) Position of antlion in trough of pit. The white line marks the pit edge. 17 18 stress where a mass is in equilibrium) (Singh 1976), and therefore will affect the angle of repose. Marachi et al. (1972) have shown that 0 decreases with increasing particle size, but they note that other studies suggest either no relationship or an opposite one. However, the actual physical factors that create the negative correlation between slope and particle size are unimportant. I have observed antlions constructing pits in several types of soil and this correlation held in each case. Thus, a pit will tend to be lined with fine sand through the differential response of particle size and 0' alone. Particle Trajectory: Stoke 's Law If an antlion constructs a pit in sand consisting of a variety of particle sizes, rings are formed around the pit in order of increasing particle diameter (Fig. 1.2). This indicates that larger particles are thrown farther during pit construction than smaller particles. Thus, in addition to a differential sand sorting on the furrow walls due to the angle of repose of sand (0'), there appears to be sorting according to the size of the thrown sand. To understand the basis of the latter sorting, the physical processes affecting sand particle trajectory must be understood. The trajectory of a particle with a given initial velocity is affected by the drag force imparted on it by friction due to air. As derived below, the drag force on a particle will vary with particle radius. The smaller the particle, the higher the drag force due to air relative to its momentum and the shorter the distance it vill travel. At Reynolds numbers below 0.1, Stoke's Law defines this force (F ) 19 (Bird et al. I96O): F = u R^(0.5pV^)(24/Re)=6 ^uRV, (I) where Re=Reynolds number=(2RVp)/u, E=particle radius, p=fluid density=0.00123 g/cm^ for air at about 25°C and 56°/o relative humidity, V=particle velocity, u=fluid viscosity=0.000184 g/cm s. The relationship between particle trajectory and the characteristics of particles can be more easily analyzed if Stoke 's Law is expressed in terms of its effect on the distance a particle travels. Here distance (D(x)) is defined as the total horizontal distance a particle travels (see Lucas 1982 for the derivation): V ^sin29 K'V sin9 D(x)= -2 [i_(4/3)-.2____] (2) g gR where K'=(7.16x10"^)c, 0=initial trajectory angle, g=acceleration due to gravity=980 cm/s , V =initial particle velocity, c=dimensionless coefficient=4.5 for a sphere (Bird et al. 1960). 20 Equation (2) is a standard Newtonian ballistics equation which incorporates momentum loss due to drag. According to Stoke 's Law, the variables that affect the distance a particle travels are initial velocity, trajectory angle, and particle radius. This equation predicts the following relationships: (l) Distance (D(x)) will increase monotonically with increasing initial velocity (V ). (2) The effect of trajectory angle (9) on D(x) will vary with sand particle size (r). The trajectory' angle at which distance is maximized will decrease from 45° as particles become smaller. As sand particle size increases, D(x) is maximal at ©=45 • (3) As sand particle size increases, distance should increase monotonically. At intermediate Reynold's numbers (2 O DC < LU 3 (D ^^ O H oo o ^ f^ -7 oo ' _i 0) Q Z '*— ' Q. < E P X I. O 0) 1- < LU 2 > _l 1- ro _J — < 1- > 3: DC >- 0 LU in DC L. Q. o DC T3 LU O 1- Z E o Q LU Q. f- TO 1- O CJ LU Q. — ' X CN LU Ol "7y 46 c) pursuit cost, d) search cost; 2) since prey density should be negatively correlated with search time, feeding time and percent consumption should decrease with increasing prey density; 3) there should be a negative correlation between feeding time (and percent consumption) and the extraction rate, since the extraction rate is inversely proportional to C. Cook and Cockrell (1978; also see Parker and Stuart 1976) also suggested that the predator should respond to the mean encounter rate and not to each individual inter-arrival interval. This implies that there should be no correlation between inter-prey interval and feeding time if the overall rate of prey encounter remains constant. These predictions are based on the fact that the predator should weigh any benefit derived from ingestion against the benefit associated with dropping the prey and searching for another. As the rate of extraction decreases, there should be a point where searching for the next prey will be more beneficial than continuing to feed on the present prey. The optimal solution to this tradeoff between ingestion and search is expressed in equation 4. Ambush Predators For a searching predator, the inter-arrival interval is dependent on search time, and therefore, the forager has some control over prey 47 encounter rate. For an ambush predator, prey arrive independently of the behavior of the ambusher. This means that prey inter-arrival intervals for ambush predators are not the same as those for searching predators. As a result, there is no tradeoff between search time and feeding time for ambush predators, and therefore, the Marginal Value Theorem (MVT) is not an appropriate model for this mode of predation. Inter-prey interval may affect prey consumption, but for reasons unrelated to the KVT. This is demonstrated by the models listed below. For the first set of models, I assume that each prey encountered is captured, and that the gross benefit derived from each prey per unit feeding time can be expressed as in equation 1 . The model for ambush predators generates different predictions at different prey densities. I will address each of these prey density regions (high, medium and low prey density) separately. Ambush model — high density At high prey densities I assume that prey are continuously available to the predator, such that as soon as the predator drops one prey, it can immediately begin pursuit of a second prey. At these densities, the predator will always be in either the pursuit or feeding phase of predation. At satiation, the predator may also exhibit some digestive pause (see Johnson et al. 1975; Sandness and McKurtry 1970). If the predator is not yet satiated, then the optimal feeding time will be similar to that predicted by the Marginal Value Theorem, except there is no search. Thus the net rate of benefit accumulation is -VOH!; 48 E g(t ) - C t - C t _ e p p e e T e P _ _ P p e e max e (c - t )(t + t ) (t - t ) (t + t ) Sep e p ® P , (5) where C is the cost of feeding per unit feeding time (t ) and the e total cost of eating is a linear function of eating time. The optimal feeding time is (see Fig. 2.2) -Dt C-[g t C(DC+g -Dt )]°*^ Dt -g p max where D=C -C P e The predictions from equation 6 are as follows: 1 ) there should be a positive correlation between feeding time (and percent consumption) and the following parameters: a) pursuit time, b) pursuit cost; 2) prey density should have no effect on either handling time or percent consumption, unless the time or cost of pursuit is affected by prey density (for example see Treherne and Foster 1981); 3) there should be a negative correlation between feeding 49 AMBUvSH/HIGHEST DENSITY MODEL Ll. LU Z LU LAG TIME HANDLING TIME Figure 2.2. Graphic method for solving the Ambush-Predator Model with high prey density. C =0 and C =0 for this qraph. e p ^ 50 time (and percent consumption) and the following parameters: a) extraction rate (see predictions from eq. 4), b; cost of eating. The digestive pause may have a varietj^ of effects on foraging, depending on how the pause constrains foraging. For example, the predator may not return to the prey after satiation, in which case the gut clearance rate and gut size will set an upper bound on feeding time and percent consumption (as shown by Holling 1966 and Johnson et al. 1975). I will model the simplest case here, where the predator can return to the prey (as shown in mites by Sandness and McMurtry 1970). In this case, equation 5 is expanded to include the cost of the digestive pause (C ) and the time required for the pause (t ) ° d ^ s^^J - ^J - c^t, - C t ^ e P_2_ ^ ^ ® s ■^rn t + t + t, T e p d (7) The optimal feeding time is \op=^^^max^V'd)^^d^d^Vp^/^max^°*' ' ^^^ The predictions from eq. 8 are the same as those from eq. 6. In addition, increases in t, and C^ should increase te . da op 51 Ambush model — low density I define low prey density as densities at which the probability of encountering a prey during either the lag or ingestive phases is essentially zero. Here the inter-prey interval is long, but this interval cannot be treated as it was with the MVT. Prey arrive at given intervals of time, regardless of how the predator uses that time. With the MVT, prey arrive at given intervals of search time only. Thus, the inter-prey interval is influenced by the amount of time the predator invests in each phase of predation. For ambush predators, prey arrive at given intervals of total time. At low densities, the sum of the pursuit time (t ), feeding time (t^) and waiting time (t =time from the end of feeding until the next prey encounter) is constant (T) and not a function of t . Here e T=t +t +t . pew I will also treat pursuit time as a constant. If waiting costs (C ) and feeding costs (C ) are negligable, then the predator should hold on to its prey until it is entirely consumed. Unfortunately, the Kichaelis-Menten function asymctotes to g at max infinity, thus assuming that the predator can always extract more from the prey. If feeding costs are non-negligible, then the predator should retain the prey approximately until the net rate of benefit accumulation drops to zero (Fig. 2.3). If feeding cost is a linear function of handling time, then the total benefit accumulated per unit foraging time is k^- "^^ >, . Q) QJ U JJ &. tw 3 0 o ^ e c ^ •H 4J U •H 0 3 c D iH U^ QJ T3 )-l O TO 5: (1) C Vj •H 0 t— i J-i « C3 -o CD tfl ^J •H CU 1 iJ x: E CO 0 3 a J3 e c < 0 •r-l 0) J-> ,c cn J-l 0) 6C <4-i C 0 •H c 1 — 1 0 Cfl 3 -u 0 0 tc X 4-1 (-1 0 -3 M-i C CO -o 0 J.) J= c 4-1 cfl (1) j-i 6 CO c u 0 •H 0 ^ 0. CO CO S-i CO c •H u m CN • 3 CO OC C •H 0) 53 UJ Q O >- H CO z LU D I- co LU o CO D lld3N3a 54 E g(t )-C t -C t -C (T-t -t ) e eeppw ep T T Thus, E gt Ct etc (T-t -t ) = _?5f_£ _f_£_ _ _P_P_ _ ■** ^ V (q) T^ T (C + t^) T T T Here c(c -c )-[(c -c )g c]°-5 t = __-5_-l^_-L_£_.^_!!?52^___ . (10) ^°P - (C -C ) e w Equation 10 predicts that the predator should handle the prey until its rate of net benefit accumulation drops to C . At handling times greater than this, it will be more costly to feed on the prey than it would be to drop it and wait for the next prey to come along. Further predictions from equation 10 are as follows: 1 ) feeding time and percent consumption should be positively correlated with g and C : max w 2) feeding time and percent consumption should be negatively correlated with C and C ; 5) prey density should have no effect on either feeding time or percent consumption; 4) neither pursuit time nor pursuit costs should have any effect on feeding time or percent consumption. 55 Ambush model — medium density At densities intermediate to the low and high density cases, prey arrive at intervals short enough to overlap with the pursuit or feeding phases. At medium densities, when a prey arrives, the predator can either drop the prey item it is currently eating and pursue the second prey, or ignore the second prey and continue eating the first (Fig. 2.4). We should expect the decision made by the predator to reflect the maximal net rate of benefit accumulation. At these densities, the inter-prey interval sets the feeding times. In fact, the pursuit time (t ) plus the feeding time (t ) are equal to P e the inter-capture interval. However, the predator should never hold on to a prey longer than the time predicted by the ambush/low-density model (once inter-prey interval drops below this threshold, the ambush/low-density model predicts predatory behavior). A few new terms must be defined: T =inter-prey interval=t +t , ir P e X=the number of intervals before the xth prey is encountered, Y=the number of intervals before the yth prey is encountered, G(X)=benefit per unit foraging time derived from eating every xth prey, G(Y)=benefit per unit foraging time derived from eating every yth prey. The net benefit of foraging (E ^) is net :,;! 56 >- UJ a. ql o z o o UJ CO LU CO > 111 DC Q. K o W < cc -J UJ oc Q. CO UJ > cc cc < liJ 0) -o >- 0) Q. o; 1) ■D (U E i- 0) 4-1 c C (1) 1_ O *-> ID in 3 e ID (U J3 ro «D > V) V U O u (U JC lld3N3S -a- CM L. 3 57 g t „ max e ^ , , , E^^. = - C t - C t . (11) net c + t ^ P ® ^ e The net benefit per unit foraging time will be g (XT^^-t ) C t C t G(X) = -H?5^-JL_F.__ . __P_P__ . __£_£__ . (12) (C+XT-tp)(XT^p) XTjp XTjp If X (12) Equation 6 generates the optimal feeding time for an ambush predator at high densities. This optimal feeding time will also correspond to the "optimal" inter-prey interval. Since G(X) decreases monotonically as the inter-prey interval increases above this optimum (see Fig. 2.5), the following predictions can be made: 1 ) at intervals larger than the "optimal" inter-prey interval, each prey encountered should be pursued; 2) at intervals less than the optimal, the best interval will depend on the characteristics of the curve from eq. 11. Prediction 1 generates two other predictions that are relevant: 5) as prey density increases, handling time and percent consumption will decrease until the "optimal" prej' ^ (>0 w 0) J3 c >»• g •H •H -O OC •a 4J ^ CD >. CD 3 J= 4-1 4-) a 11 •H O 11 * W XJ H -K c 0) -a Q c • • CO CO .-^ s 0) 0) o /-^ •H TJ ^ /— s 4J o *^^ X ■H 2 j: v.^ Cfi CD o c >. 3 OJ 4J X T3 'H £ c tc < o >, c •H OJ 0) 0) 4J 1-1 o X CO D. 4-1 I-l x: D 0) M c e 4-1 -H 01 3 CO ffi X a •H ~-^ 4-1 • o T3 J C CO Q) CC •K O 6 3 * •H u ^- X) ^ 4-1 •H , M E Q) XJ •H XI 4-1 CO -O ^ 03 U_l 0) t^ cn O CJO 4J Q) •H c CJ TD X (U •H -H O 4-1 4-1 60 13 2 CO CO 0) I- >-i V4 U >^ 0 o c 4-t U-, 4-1 u-^ ■H 0) QJ cn o z 1^ 6 c II , O -H 0) •U 4-1 c o' • to U-1 T3 OC 3 • 0) C Q • CN U -H ^ »- D,--! "^■^ o 0) T3 X •H ^ CD C tn > 2 CO 3 CO 6C ) sj X X X •H O 11 E a p:- iw * < X' 59 aau Q- -O (X)D v.'"^t' 60 interval is reached; 4) since each prey item is pursued in this region, variation in prey encounter should be correlated with variation in handling time and percent consumption. Prediction 3 is identical to the predictions from the MVT; thus at medium densities the ambush predators should treat prey similar to searching predators. However, prediction 4 is different than the analogous prediction for searching predators and is identical to one of the predictions from the GLM. Capture Probability and Ambush Predation The models listed above illustrate that many factors will affect the predictions from optimal foraging. I have focused on two factors, the mode of predation (searching vs. ambush) and the effect of density on the predicted predatory behavior. In this section I develop a model for ambush predators which incorporates capture probability. Griffiths (1982) suggested that there should be selection for reduced handling time if the capture probability is lower when the predator is handling a prey than when it is "empty handed". The model presented below explicitly demonstrates this relationship. Another important factor is whether or not the predator returns to a prey item once it is dropped. For example, damselfly larvae apparently do not return to prey (Johnson et al. 1975), whereas antlion larvae may cache partially utilized pr'^y on the pit wall, capture the second prey, then return to the first (pers. observation). I will assume that the predator can return to the first prey, so that if 61 a prey comes before the predator is finished with a previously captured prey, it will cache the first prey until it can go back to it and finish it. Let A =encounter rate of prey, P^'^Pi'obability of capture while handling a prey, P =probability of capture while empty handed, P,=P -P , a 0 w N=total number of prey handled in a foraging bout lasting a given length of time, T , T,=total time predator devotes to a given prey=t +t . " p e To simplify the model, I will also assume that the cost of eating and pursuit are negligible. This assumption will not affect the qualitative predictions of the model. First the number of prey eaten while handling a prey (N ) and the w number eaten while empty handed (N ) must be calculated: N^= A (T^-NT)P^ . (14) The total number of prey eaten (N) is therefore N=K^+K^= ANTP^+ A (T^-NT)P^ . (15) Solving for K, ATP ATP ,. .....1.2..... . ...iifs.. . (,6) The benefit associated from each prey. Kt g(t ) = — 2___ , (17) C + t e 62 times the number of prey yields the gross benefit for the foraging bout; (Kt )( XT P ) B = g(t )N = ? L2____ . (18) (CnJ(uXT^P^) The benefit gained per unit foraging time is (Kt )( XP ) e 0 ^T ^^^^^^^*^Vd^ The optimal solution to equation 19 is t = (Ct + f'^ . (20) eop ' p ^p ^ ^ ^ 'd (19) From equation 20, a predator should decrease handling time as the difference in capture probability (P -P ) increases. This is because 0 w there is an added cost to holding onto a prey that must be weighed against the diminishing return from that prey. Equation 20 is unique among the ambush models presented here because it is the only model that requires that the predator "anticipate" the next prey, or at least modify its behavior before the next prey arrives. Thus, differences in capture probability should affect how the predator treats variation in the inter-prey interval. Discussion Different predators appear to exhibit a wide diversity in their responses to prey. Also, as conditions change, the behavior of a single predator may be predicted to change considerably. Many predators (for example Plethodon; Jaeger and Barnard 1 981 ) may switch from ambush to 63 searching predators as prey density fluctuates. Some predators may continuously reach satiation (ex. mantids in Rolling 1966, damselfly larvae in Johnson et al. 1975), while others may rarely if ever be satiated (ex. hummingbirds in DeBenedictis et al. 1978; antlions in Griffiths 1982). This diversity is an important consideration in using an optimal foraging approach to partial prey consumption, since predictions change both quantitatively and qualitatively with changes in predator or prey conditions (see Table 2.1). One primary focus of a number of papers to date has been the evaluation of proximate vs. optimal foraging models. The Marginal Value Theorem (Cook and Cockrell 1978; see also Sih 1980a) was originally used to show that the Gut Limitation Model was inadequate. Cook and Cockrell (1978) showed that for a cocinellid and a notonectid, percent consumption and handling time both decreased with increasing prey density (predicted by the MVT and GLM) and that individual feeding times were independent of the previous inter-catch interval (predicted by the MVT, but not by the GLM). Giller (1980) repeated the experiment on notonectids and found that individual feeding times were not independent of the previous inter-catch interval (predicted by the GLM). Giller (1980) also found that handling time per item decreased through the foraging bout independent of prey density, suggesting that the predator may be forming a search image through some optimal feeding mechanism. Griffiths (1982) proposed the Digestion-Rate-Limitation (DRL) model to explain this decrease in handling time in notonectids and showed that the DRL Model applied to antlion larvae as well. He also showed that ant lion larvae fed at low feeding rates do not change the percent consumption with changing feeding rates, as predicted by the GLM. 64 Table 2.1. The effect of predator and prey characteristics on predictions from partial prey consumption models. '+' = positive correlation, '-' = negative correlation, '0' = no correlation expected, 'N/A' = not applicable. handling time/percent consumption variable GLM DRL MVT A/H A/L A/M* A/CP prey density a) near satiation -/- -/- -/- 0/0 0/0 -/- _/_ b) no satiation 0/0 -/o -/- 0/0 0/0 -/- -/- intercapture interval a) near satiation +/+ 0/0 N/A 0/0 V+ 0/0 b) no satiation 0/0 0/0 N/A 0/0 V- 0/0 cost of pursuit (C ) 0/0 0/0 v+ +/+ 0/0 0/0 cost of search (C ) s 0/0 0/0 -/- N/A N/A N/A N/A cost of eating (C ) 0/0 0/0 -/- -/- -/- 0/0 cost of waiting (C ) w N/A N/A N/A N/A -/- N/A pursuit time (t ) a) near satiation -h + /+ */+ ^h 0/0 0/0 V+ b) no satiation 0/0 0/0 V+ ^h 0/0 0/0 +/+ extraction coefficient (c) a) near satiation -/- -/- -/- -/- -/- 0/0 -/- b) no satiation 0/0 0/0 -/- -/- -/- 0/0 -/- * predictions for interprey intervals greater than the 'optimal' inter-prey interval only. The predictions for no correlations are due to the fact that all prey should be pursued (see text). 65 From the models presented in this paper, it appears that the arguments over proximate and optimal foraging mechanisms in notonectids addressed the wrong optimal foraging models, since Rotonecta is an ambush predator (Gittelman 1974). Giller's (1980) results are predicted hy both the optimal foraging model for ambush predators and the GLM. The lack of change in percent consumption for antlions (Griffiths 1982) is also predicted by the ambush optimality model. Thus, the differences between proximate models and the correct optimal foraging models are non-existent for the parameters addressed in the literature cited above. Griffiths (1982) also suggested that in many cases proximate and optimal models will generate similar predictions, though he incorrectly equated the predictions from the KVT (which was the incorrect model anyway) and the proximate models (DRL and GLM) at low prey densities. However, it seems counterproductive to compare the two sets of models in the first place, since the goals of the different approaches are dissimilar. Rolling's (1956) goal in modeling proximate mechanisms of predation was to generate a realistic model that could be used in a number of theoretical studies. These studies include an analysis of functional and numerical responses, and the relative advantages of digestion rate, prey size or predator size. He also suggested that his model could be used to test whether the mode of predation exhibited by a predator maximized energy input or minimized energy output. Thus, his proximate models required a complete knowledge of predatory behavior, but could then be used to test other aspects of predation. On the other hand, optimal foraging models attempt to predict the behavior that should be expected from an organism based on our knowledge of the factors (or currencies, Pyke et al. 1977) that may be important in the life of that k-'' 66 organism. The output of these models says nothing about the proximate mechanisms that drive these behaviors. It is implied that the evolution of proximate mechanisms should proceed in such a manner as to approximate the optimal behavior patterns. The models are used to test how well our understanding of the important factors account for the evolution of the behavior (Maynard Smith 1978), irrespective of the exact evolutionary pathway that culminated in the behavior. All optimal foraging models rely on a set of assumptions. For example, optimal foraging models have all assumed that the extraction rate curve is constant. However, the DRL model proposes that the curve may change with feeding rate. This change does not refute the optimality approach, it simply requires a change in the assumptions about the rate curves. In fact, an increase in extraction rate with increasing prey densities undoubtedly will increase the net rate of ingestion over the entire foraging bout. Thus predators that can increase extraction rates will probably do better than predators whose rates remain constant. In a review of optimization theory, Maynard Smith (1978) said that biologists need simple biological models that hold qualitatively in a number of cases, even if they are contradicted in detail in all cases. He implied that a qualitative fit to predictions will generally bring the researcher closer to an understanding of the problem in question. Unfortunately, generalizations can lead us to accepting models prematurely. This problem is aptly demonstrated by this review of models about partial prey consumption. In a sense, part of the question concerns the definition of detail. For example, one could argue that the expected correlation between intercapture interval (given a constant 67 density) and handling time is irrelevant detail, in which case the difference between some of the models presented here is unimportant. However, I would argue that one of the strengths of optimization theory is that a quantitative prediction can be explicitly generated and tested. A number of factors can contribute to the lack of quantitative fit to a model. Three of the most important of these are constraints on foraging behavior (including both physiological constraints and ecological constraints), the failure to include important parameters into the optimization model and the divergence from an optimal solution using a satisficing criterion (see Simon 1955). The lack of fit to an optimization model is bound to yield a greater understanding of the system when these alternative factors are pursued. But this is a reasonable pursuit only if models specifically suited to the system are tested. CHAPTER III PARTIAL PREY CONSUMPTION BY ANTLION LARVAE Introduction In chapter 2 I addressed existing models of partial prey consumption and compared two different types, mechanistic and optimality models. The models were found to generate different predictions under different conditions. Thus, although some generalizations may be made concerning partial prey consumption, even qualitative predictions cannot be formulated without restricting them to a specific system. This chapter is a test of Griffith's (1982) "digestion rate limitation" model and the optimality models from chapter 2, using antlion larvae as predators. Antlions are particularly appropriate for testing the models since Griffith's mechanistic model was derived with antlions in mind. I will first present the predictions and tests of Griffith's model. I then derive and test predictions of an optimality model appropriate for the antlion system. Antlions construct conical pitfall traps in sand that aid in the capture of arthropod prey. Once a prey item is captured, the antlion injects digestive enzymes into the prey and ingests the predigested material (Wheeler 1930). The exoskeleton is never eaten, and therefore the antlion never consumes the entire prey. As I show below, an antlion may also discard a prey before all of the extractable prey biomass is ingested. 68 69 Digestion Rate Limitation (DP.L) Model; Predictions Predictions Griffiths (1982) showed that the rate at which an antlion ingests prey increases as prey-capture rate increases. This is presumably due to the fact that antlions produce digestive enzymes at a higher rate when prey capture rate increases. Handling time was shown to decrease with increasing capture rate (Griffiths 1982), which is consistent with this model. Griffiths also suggested that the proportion of each prey extracted should not change if prey are not simultaneously encountered. He predicts that at relatively low feeding rates, antlions should simply extract all they can from their prey irrespective of encounter rate. The prediction, which originated from the work of Holling (1966), is that partial prey consumption is caused by the filling of the gut. At low prey densities, the gut of the antlion will never be full (if the prey is small enough, as will be true in this experiment). Thus partial prey consumption should be independent of prey density at low feeding rates. Methods To see whether antlions followed the two simple predictions generated by Griffith's model, we fed fruit flies (Drosophila melanogaster, vestigial winged) to antlions (third instar Myrmeleon mobilis; identified according to Lucas and Stange 1981) at four Q different feeding rates. Antlions were kept in the lab at 24 C for at ■y 70 least seven days prior to feeding and fed one fruit fly per day during this acclimation period. The larvae were then fed one pre-weighed (to +/- 0.00001 gm) fly per day (FS-1 ) for 3 to 5 days. For each fly, total handling time was m.easured and the carcass was weighed immediately after it was discarded by the antlion. The difference between the initial weight and final weight was calculated as the extracted wet weight. Percent wet weight extracted (predicted to be constant) was calculated by dividing the extracted wet weight by the intial wet weight. Antlions were then divided into one of three groups corresponding to the remaining three feeding categories: FS-8 (1 fly per 3 hr), FS-24 (l fly per hr), or YS-A8 (1 fly per 0.5 hr). For FS-8, antlions were fed from 4 to 7 fruit flies in a row; for FS-24 they were fed from 6 to 17, and for FS-48 they were fed from 5 to 12 in a row. Each run (FS-1 then FS-8, FS-24, or FS-48) was made with a different antlion. A pilot study suggested that antlions take, on average, less than 30 min to handle a fruit fly. Thus, the maximal feeding rate (FS-48) was set at a rate just low enough to ensure that an antlion never encountered a fly before it had finished the fly it was handling. Thus, by definition (see Chapter 2), the antlion was feeding under low density conditions. Multiple linear regression (GLM procedure in Barr et al. 1979) was used to test the predictions of the DRL-model. Since handling time decreases with increasing encounter rate (Griffiths 1982), this should result in a negative regression coefficient on handling time when regressed against feeding rate. This v.-as tested to determine whether our species forages in the same manner as Macroleon quiquemaculatus, the species studied by Griffiths (1982). We were unable to control for variance in two factors: initial fly weight and individual variation 71 among antlions. The antlion effect is also compounded by the unequal number of flies given to each antlion. In an attempt to account for variance associated with initial fly weight and inter- individual effects, these two factors were added to the regression model. Thus the regression model used to test the hypotheses was T^ = b^ + b^FS + b^I + b AN + e , where T = handling time, FS= feeding schedule, I = initial fly weight, AN = antlion "name", e = error term, b -b^ = regression coefficients. 0 :) This model allowed us to test for the effects of feeding schedule on handling time, independent of the initial fly weight and antlion differences. This same model was used to test for the effects of FS on percent of each fly consumed (predicted to be constant) and total extraction rate (predicted to increase with increasing feeding rate). Percent consumption was transformed using the arcsin square root transform (Sokal and Rohlf 1969). Antlion name was treated as a class variable (see Barr et al. 1979), because it was nominal scale data. Antlion weight was substituted for antlion number in the above regression equation to determine if larger antlions could handle flies 72 more efficiently. This information was used to tuild the models listed below under Optimality Model. Results As predicted, handling time decreased with an increase in feeding rate (Table 3-1). Contrary to predictions, the percent of each fly consumed dropped significantly with an increase in feeding rate (Table 3.1). Thus, antlions discard prey before they are totally empty even under low density conditions. As prey capture rate increases, the antlion extracts less from the fly, even though it never encounters two at the same time. Thus, the predictions of Griffith's model are not supported. Antlions appear to be regulating prey handling behavior at a finer level than that predicted by the mechanistic models listed in Chapter II. Below I test an optimality model that I derived to test whether antlion feeding behavior was consistent with the prediction that they were maximizing net energy during handling time. The optimality model should not be treated as an alternative to a mechanistic model, since they address different aspects of the same behavior. They appear as alternatives here because they have (unfortunately) been treated as such in the literature. Deterministic Optimality Model In chapter 2, I derived an optimality mod il of partial prey consumption under low prey density conditions. The model predicted that 73 Table 3-1. Linear Regression statistics for handling time and extraction of biomass from fruit flies. Table A includes antlion name as an independent variable. In Table E, antlion weight was substituted for name (see text). Each number is the probability that the regression coefficient associated with the independent variable is zero. The sign on the probability is the sign of the regression coefficient. AS % ext = arcsin square root transform of percent wet weight extracted from fruit fly. FS = feeding rate, initwt = initial fruit fly weight, lionnm = antlion 'name', lionwt = antlion wet weight, extrate = extraction rate. A) INDEPENDENT VARIABLES dependent 2 r 0.26 model total P>F for variable FS initwt lionnm* .0001 df 79 df 552 F 2.1 m.odel AS % ext -.0051 +.0254 .0001 \ -.0001 +.0001 .0001 0.74 79 552 17.2 .0001 extrate +.0001 +.0001 .0001 0.64 79 552 10.7 .0001 * Class variable, no slope estimatable B) INDEPENDENT VARIABLES dependant ^ model total P>F for variable FS initwt lionwt r^ df df F model AS % ext -.0234 +.0020 -.2898 .03 3 552 6.0 .0006 ^h -.0001 +.0001 -.0015 .46 3 552 156.3 .0001 extrate +.0001 +.0001 +.0020 .30 3 552 80.1 .0001 74 ■ >• percent consumption should not vary with prey density. However, the model assumed that the extraction rate curve (ie. extraction rate as a function of handling time) remained constant with changes in prey capture rate; this assumption is clearly violated here. Thus, predictions concerning optimal prey utilization can only be generated after the prey extraction rate curves are constructed. All optimality models are couched in terms of a currency or currencies (Pyke et al. 1977). Energy is by far the most common currency, although others have been used (see Pulliam 1975; Rapport 1980; Greenstone 1979; Belovsky 1981; Westoby 1978; Owen-Smith and Novellie 1982). I use energy as a currency here for two reasons: (l ) It is the most likely currency with which to estimate foraging costs to the predator. Of course, if there is no cost to the animal in terms of any currency, the forager should simply eat the whole prey or extract as much biomass as it is capable of extracting. (2) At eclosion, antlion adults weigh 50 percent of their pre-pupal weight (Lucas, unpubl. data). Thus the weight of the larva at pupation will determine the weight of the adult. If adult weight correlates with fitness in antlions (as it does in many other insects, see Schoener 1971), an increase in the net rate of energy intake as a larvae should increase the fitness of the adult. There are a number of variables that must be incorporated into a model used to predict optimal prey utilization. These include: (l) an extraction rate curve (here expressed as wet weight per min handling time), (2) the conversion of extracted weight to calories, and (3) an expression of the energetic cost of foraging. The extraction rate curve , j- must predict the biomass extracted at any given time during the handling {^.^ 75 of a prey. From these three variables, the net energy intake (gross energy minus the cost of extraction) can be calculated as a function of handling time. From this function, the handling time that maxim.izes net energy can be calculated. This optimal handling time can be compared to the observed handling time to determine whether the criteria on which the model is based, are good predictors of the foraging behavior. Thus, the three variables above are descriptive models of foraging. These models are then combined into a predictive model of optimal behavior. Methods (1 ) Extraction rate curves Wet weight extraction curves were constructed for each feeding schedule. Antlions were fed flies that were then taken from the antlion after 2, 5, 10 or 15 minutes. These data were combined with extraction data for uninterrupted feeding times (Figs. 3.1-3.4) to generate the extraction rate curves. From Table 3.1 we know that antlion weight and initial fly weight affect the extraction rate. Initial fly weight additionally will influence the percent of each fly consumed. The construction of the extraction rate model was based on these relationships. I fit the data to three types of curves: (l ) a Michaelis-Kenten function, (2) an exponential function, and (3) the power function listed below. The third model proved to be the best predictor of the data, and was therefore chosen as the extraction rate curve for the optimality model: -I G.GCIG- ^ i i 1 0. u002- ] 4 O.GCOl-^ i 1 J o.ooooJ 76 :czn<^^'r.= i 3 „ □ ^ 2 ^ :=-^ c n ~ c E □ C n a O.OOOS-^ ff ^ 0.0005 J 0° i n r- , o.aoo7-i ^ A ^ ^ X '^ ^ ^ 0.0005 J □ J n 3 a •^ A -^ AA - "^ A q ^-.AAA't^A'^A A > _^A i A A -.- -1-4- -1- + ++■ + A + -t-* J- .^ ^ ^ - S 0 0.0004-3 Q S * ■ j + O.OOOoJ n 0 3 6 9 12 15 le 21 2u 27 30 33 35 35 ^g yg yg TIME Figure 3.1. Biomass extracted from fruit flies (extrwt) at various handling times (time) when the feeding rate was one fly per day. Extracted weight is expressed in grams wet weight, time in minutes. The symbols represent different initial fly weights: cross: initial weight -rO.OOO? gm; triangle: 0.0007 gm (initial weight < 0. 0009 gm; square: initial weight > 0. 0009 gm. 77 EXTRWT 0. 0013-! "^EnsrH=3 0.0012- 0.001 H J 0.0010 J 3 o.ooosJ ■i 4 I 1 o.oooeJ 0.0007 0.0006-1 0.0005-i J -i o.oooyJ 0.0003-1 0.0002-1 0.0001-i D G n CD A D D .r A A .A ^'^ A 4- A a"' + + * + -I- a. 0.000G-: I ' ' ' ' 0 10 15 20 TIME 25 30 35 Figure 3.2. Biomass extracted from fruit flies (extrwt) at various handling times (time) when the feeding rate was eight flies per day. Symbols as in Fig. 3.1. 1 J J 0, 001 1- FFEDSCM=24 c D 3 0.0010-3 oos-3 0.0 0.00C8J -1 0.0007-q 0.0006^ -i 0.0005- ° D, ° CC ^ _ = A □ A i □ ° ^ "^A K. A A ' A A ^ A^ A-^ A ° -IS ^AA ^ £ A ^A A A A "^AA A *''+ ^ ^ A ; \ 4- 4- 0.0004- 0.0003J 0.0002-; B A 0.0001- 0.0000- LO 15 20 TIME 25 30 3= Figure 3.3. Biomass extracted from fruit flies (extrvt) at various handling times (time) when the feeding rate was 24 flies per dav. Symbols as in Fig. 3.1. 79 E^''°tv* '^EEDSCH = 48 O.OOll- °r. i -I O.GOIO-: H 0.0009- 0.0008^ 0.0007J 3 0.0006- 0,0005-^ o.ooom- □ ^; □ n AA ^ A A A A . A AA A A + A A 0.0003J 3 0.0002- 0.0001- O.OOOG- 10 20 TIME 25 Figure 3.4. Biomass extracted from fruit flies (extrwt) at various handling times (time) when the feeding rate was 48 flies per day. Symbols as in Fig. 3.1. 80 k k k Gg(T) = XKk^-k^I ')(l-(k^-k^I ^k^W 8)- = ^T-L)) ^ ^^) where G (t) = the gross gain, in terms of wet weight, extracted from the fly after time T, T = time starting from introduction of fly and ending at the release of the fly, I = fly initial wet weight, W = antlion wet weight, L = lag time from time of introduction of fly to time when the antlion first begins to extract biomass from the fly, c, k, -ko = constants, 1 o X = conversion from wet weight to calories, see below. The coefficients were estimated using a non-linear least squares method (program NLIN in Barr et al. 1979). (2) Wet weight to calorie conversion Dry weight was measured on 25 fruit flies that had been weighed wet then dried for 5 days at 60 C. The regression equation fit to these data was used to convert wet weight to dry weight in the extraction rate curves. Drj' weight was converted to calories using the following conversion for fruit flies (Cummins 1967; Jaeger and Barnard 1931 ): 1 gm dry weight = 5797 cal. v.^ 81 (3) Metabolic cost Both eating costs and waiting costs were measured as rate of oxygen consumption using a Gilson respirometer. All measurements were made at a temperature of 24 C. The respirometer was allowed to equilibrate for one hr before readings were taken. Two cm of sterile sand was placed at the bottom of 15 ml Gilson flasks into which antlions were introduced. Waiting costs were measured after antlions constructed pits in the flasks. Readings were taken every 30 min for 2 hr for each antlion. The mean oxygen consumption rate from these data was used as the waiting cost for the antlion. Measurements were used only if the antlion did not move during the 2 hr measurement period. Eating costs were measured by placing a live fly in a small open-topped vial glued to the inside of the Gilson flask. The vial was sealed on top with a steel ball then the sand and antlion were placed in the flask. When the respirometer had equilibrated, the steel ball was lifted up with a magnet. The fly would then jump out of the vial and fall into the pit. Readings were begun after 2 min, which was enough time for the fly to be killed by the antlion. I was unable to measure pursuit+pre-ingestion costs. This cost was assumed to approximate C , since the level of activity was similar for the two behaviors. Errors in making this assumption will not be very major since the calculation of the optimal handling time is independent of C (see below). All antlions were kept in the lab for two weeks and fed two flies per day before calculating metabolic rates. Each value represented a single individual. To calculate an analytical expression of metabolic rate for the two behaviors, the classic metabolic rate equation (cf. Kemmingsen I960): 82 MR = aW^ , where W = weight, and a,b = constants, was expanded to include a covariance term between waiting and eating. By adding both terms to a single model, T am assuming that there is a linear relationship between the log transformed metabolic rates for activity (eating) and waiting. This is true of two other species of antlions (M^ crudelis and M^ carolinus) for which I have a large data set of metabolic rates (unpubl. data). The addition of both eating and waiting costs increased the degrees of freedom of the model, which should decrease the error in the estimates of each coefficient. The expression is given here in the linear form used to fit the regression coefficients: In(MR) = ln(a) + b ln(w) + cE + dE ln(w) , (3) where E = 1 for an antlion eating and 0 if the antlion was "waiting" and a,b,c,d = constants. The constants were fitted using the least squares technique (program GLM in Barr et al. 1979). Metabolic rate- was converted to calories by assuming an R.Q. of 0.8 (which is the most reasonable R.Q. for an insect of this type; K. Prestwich, pers. comm. 1982). Thus one ul 0 is equivalent to 0.0048 cal (DeJours 1975). i 4<- "- 83 (4) Net extraction rate curve Net energetic gain was derived by subtracting the energetic cost of eating, pursuit + pre-ingestion, and 'waiting' (the non-foraging time spent between prey handling and pursuit times). Thus net gain, G (T), is G^(T) = G^(T)-CpL-C^(T-L)-C^(T^-T) , (2) where C = energetic cost of pursuit + pre-ingestion behavior, C = energetic cost of eating, C = cost of waiting, and w T = total time = T + waiting time. (5) Derivation of Optimal Handling Time Here I derive the optimal handling time predicted from equation 2. The derivation assumes that prey arrive at fixed intervals and that the extraction rate curves can be approximated by the deterministic model given bj' equation 1 . The constants from equation 1 can be combined as a shorthand: k k k \ G (T) = I(k^+k21 ^)(l-(k^+k^I ^+k„W S)-c(T-L)^ = k (1-k, -^^-->). (4) a D Thus G (T) = k (1-k -"^^-^^) - n (T-L) - C (T^-L) - C L , where C , = C -C . a e w W T The optimal handling time is defined as the point where the partial rate of change of G^(T)/T^ as a function of handling time is zero (cf. Charnov 1976). Thus the optimal handling time is where d G(T)/T^ -k^k^-"^^-^)(ln(k^))(-c) C^ 9 (T-L) T^ " " 't^ " ° ' Replacing the left hand side of equation 5 with zero and rearranging we get C, -c(T-L) — = k ln(k, ) Ck ^ a Rearranging and taking the natural log of both sides C ln( ) = -c(T-L) ln(k ). ck ln(k, ) a b From the above equation, the optimal handling time (pursuit + eating time = T ) is op C, .ln(— J ) ck ln(k, ) T 5...J , . (6) ^ ln(k^)c '^4 85 The optimal extraction weight and optimal percent extraction can be estimated by replacing T with T in the equation for gross extraction op weight, G (t) (equation 1). Thus both the optimal (predicted) handling time and the optimal (predicted) percent prey consumption can be found using equation 1 and 6. Once the variables listed in equation 2 are measured, the foraging response of tre antlion in terms of handling time and percent utilisation can be compared to the predictions from the models. Results 1 ) Extraction Rate Equations The gross gain model, G (t), produced a good fit to the data when initial fly weight and antlion weight were included as variables in the model. For all feeding schedules, over 99 percent of the variance was accounted for by the model (Table '5-2). As Griffiths (1982) had found, extraction rate generally increased with increasing feeding rate (Fig. 3.5). There was a simple linear relationship between fly wet weight and fly dry weight. Therefore wet weight can be multiplied by a constant to convert to calories (Table 3.5). Estimates for eating and waiting costs are also listed in Table 3.3. . ;) a e •H tot •T3 0) a> S-j o D. O O m . 0) o > u en 3 TO u 3 ij /-, 60 i O. --H 14-1 Cfl 0) .. •H e 1^ OC O .-H 3 u 3 -U O J= oc (U -H 4-1 (U n S V4 c c o O -rt • •H i-H 4J 4J 4J X CJ C 0/ CO < 4J ^ 4-1 c K! • •H Qj ^-v 4-1 -a T2 X 0) 0) (1) 4-1 4J J-1 cn U •H •H 0) rH T3 0) (1) tn en ^ ^-- u &H 0) 1—1 4-1 j cr Vj 3 Q> , 4^ fc XI o ^^ «e'«'* 87 ■ CD 00 00 • • • • o 00 r — (\.i § o3 C\) OvJ • • • » CM r<~N t^ OvJ O O o O • • • • o cr> •H o •H -P O CD +> X 0) 3 o f1 ja -p •H Ch o e u CO a> c •H rH I c o CM -P . > t^ t<^. ^ O '- K\ if\ ^— VD • a> o^ 00 o CM Lr> o • 1 1 ^ CM "t VD 00 q — ro VO CJ^ O o •^ ^ • • CT^ • »- "— • LP. ir\ '^ o O O ^— ^— ^— q — X X X VD , — 00 [> cr> ■^ CT-i o • c^ 00 CM f~ • • ' 1 T *- LTn ir\ ^f^ CM • • t 1^ ^ — CM o t<~N ^ — ir\ LP • ,— q — CM ■^ *" ■^ Od o CO cr. ^^ OJ !>■ u: "x CM t^. rvi CO o X X X [> CO cr. t<-^ •^ c^ • • • ^ •-- cr , , CM o CM '^ >^ CM CM [^ ^ CO cr. OO CO' CO • • • • '^ CO '— cc CM 'd- 89 Table 3- 3- Fruit fly wet-weight/dry-weight regression equation and metabolic cost equations for eating and waiting behavior in antlion larvae. Vet weight-dry weight regression ln(dry weight) = -0.63 + 1.00 ln(wet weight) dry weight = 0.23 (wet weight). N=25, F=35, P<.0001, r^ = 0.602. Metabolic Rate Equations In(ME) = 1.983 + .910 ln(W) - .255 E - .499 (E ln(V)). metabolism measured as ;al 0 /hr N=28 (16 eating, 12 waiting), F=144, P<0.0001, r^=.947 Caloric Cost of Eating* (assuming R.Q. =.8) eating C = 0.0043 W*'^'''' waiting C = 0.0077 W^^*^ w * cost calculated as cal/min 90 2) Verification of the extraction rate curves Before using the optimality model to predict optimal handling time, I first tested to see whether the extraction rate curves would predict observed values. If the descriptive rate curves failed to reproduce the data, the predictions from the optimality model would obviously be irrelevant. I tested this by comparing predicted values from the extraction rate curves against observed values from antlions that were allowed to feed until they discarded the carcass. This was done by comparing the percent wet weight extracted by the antlion to the percent wet weight predicted by the model for the time required to handle the fly. This difference was tested statistically by arcsin-square-root transforming the percents then subtracting the transformed values. A one-sample t-test was used to test whether this difference was significantly different from zero. Data from each feeding category were divided into four fly weight categories, since percent weight extracted is influenced by fly initial weight. In 16 comparisons (4 FS x 4 weight categories), only 2 differences were significantly different at P=0.05 level (Table 3 -A). In fact, with 16 simultaneous tests, this probability level is an over-estimation of the true alpha level. This is because, by chance, one in 20 comparisons may be expected to be significantly different using this test even if the two variables are drawn from populations that are statistically indistinguishable. Thus, the model can be treated as a fair estimator of the data. 91 Table 3«4. Optimal and observed values for handling time and percent prey consumption by antlion larvae. T^p = optimal handling time. T-u ~ observed handling time. OAS %ext = observed mean arcsin square root of percent extraction. PAS ^ext = predicted optimal arcsin square root of percent extraction, ^extop = optimal percent extraction (from PAS ^ext). ^extob = observed percent extraction (from OAS text). M-0 AS^ext = difference between the observed percent extraction (transformed) and the percent extraction predicted from the model for the observed T^. Mean fly initial weights (initwt) and antlion wet weights (lionwt) are also given. Fruit fly weight categories (WC) are as follows: 5= less than 0.0006 gm; 7=. 0006-.0008 gm; 9= .0008-. 0010 gm; 11= greater than .0010 gm. Std=standard deviation. FS/WC ?AS?ezt 0AS2ext 'extop 2eitc3b M-OAS'ext N initwt lionwt 1/5 29.7 ie.9*** 1.279 ;.22i*** .917 .88-1 -.C15ns V .OOC52 .0278 std 1.5 4.1 .001 .057 .072 .00005 .0094 U" 34.4 21.9*** ■■•230 1.23'*** .913 .889 -.CiOns 93 .00070 .0303 std 1.7 3.4 .001 .052 .054 .00005 .0118 1/9 39.3 25. 6*** 1.23' 1.233*** .916 .890 -.007ns 35 .00088 .034; std 2.2 5.1 .0004 .064 .0^5 .00006 .0139 . ,' 1 ' -ic . i std 3.2 5.3 3*. 4*** •.2S1 1.251*** .91s .901 .0002 .043 -.013' .049 43 .00110 .C35~ .00006 .0108 8/5 23.; 1S.4*» 1.205 1.156ns .872 .838 .026n3 std 0.3 3.5 .001 .070 .076 8/" 24.- 21."*** '.209 1.216ns .875 .879 -.015n3 std Co 2.7 .002 .058 .060 8/9 26.7 23.''** 1.216 1.217ns .879 .880 -.009ns std 0.7 4.3 .003 .050 .050 3 .00056 .0303 .00002 .0096 33 .0OC7O .00007 24 .00088 .00005 .0064 .0292 .0064 .066 . 8°6 .381 305 ns 352 25 .00113 .00010 .0301 .0066 24/5 24.* '5.3*** 1.192 I. 139ns .363 .861 -.052* std 1.3 2.3 .CO' .064 .079 24/" 25.3 19.2*** 1.195 1.153** .365 .846 .0C3ns std 2.0 3.4 .001 .072 .068 24/9 27.0 21.5*** '.202 •.'0"ns .270 .867 -.Ol6n3 std 2.'' 4.5 .003 .076 .072 24/' • 29.4 24. 2*** '.220 '.133* .382 .360 .012ns std 5.- 4. a .011 .085 .066 12 .00055 .0350 .00003 .0143 61 00069 00005 .0143 00033 .03=0 00005 .0151 001 1 1 .0339 OOC09 .0133 IT 92 Table 3-4 Continued. ^S/; 21.0 -6. --3 1.2-" • .Zii 3tc: :.l 3.3 .0CC3 .055 ^^oTns 1 .0CO5a .0486 061 80 •0001 .0093 % ff^f**'--^'^""' -83 .537 -.01-3 32 .OOO-O .0.05 -° -^ .X,. .06o .063 .00005 .0139 -S/9 23.? 21.1*" 1.231 1.226C3 .389 .886 -.0^3c3 '0 CCCH-^ --a<; 3- 0.0 ..o ..04 .055 .060 .00005 .0132 1(d' f^' 7^°' ^;?" \;?f °^ -^CS .298 ■ .001n3 9.00112 .0368 •056 .00010 .0090 3td 8.5 1.6 .019 .062 ns - difference v.a- significant at ?>0.05 - difference significant at 0.05 CO -•-' Vh O C 03 03 * 03 03 -CI II O 03 0 CC 03 ■T3 C C -H o s o 03 O o 03 -P -P •- C o CO cd CD -P P to 03 03 T3 J= -P -P c o •H C O II -p 03 2 G 03 > 03 Sfl ft ft I o -P X 03 CO O CO T3 03 S 03 0) ^ II +J Oh &a a K> ITS o> c . tj o 3 O -P 03 03 ft 03 f-l CO ^ O 03 CO II J= e-i O Eh ad o ■p c CO c o o 03 03 -p C) c 03 CO 03 •H 03 -P c o col 0) 00 c o C3 W 03 03 O -p O 2S c 03 ^ CO 03 1— 1 a -p -4 03 o U K en •H -oococj>xicM':j-Lrv-=a-t--LrNC~--- I c-'^r-c~-cD>xic~-ooooc^v£)v£5r--GO I I I a • • a 1 >- CM CO t^ f^ "sf o CO I I OMKMH&qKICJ>WOOEi3 I t O I ir\----if\a>c>---^o:)f<~\0 1 oouooofaooowo I I aoo I r-a^[-oO(T>c^aD^-ooco •' 1 N ^ N ' -P 1 v^ , >> • 03 03 1 VD -^ ft U 1 CO ft 1 T r 1 o 1 ^^^ "-v^ 1 03 JZ 1 + -+- 1 03 -P 1 un , — 1 -H 1 • • 1 X s 1 h^ en 1 ir\ ^ — 1 ^f^ 1 ■'""^ 1 C 1 0) s 1 • 1 e -~- 1 ''■"^ 1 O 1 •!-! CM 1 II 1 -^^ >> ft 1 03 o" 1 0) I- ^ 1 ft ft O^ ^—s 1 CO 1 v-O 1 O -P 1 ir\ 1 03 3 ^s^ (T. 1 o; o + 1 x: C^ 1 X -p • r— 1 1 -r-l o 03 1 3 03 1 03 •H 1 u Uj 1 3 , — ^ ^ — N V • 1 -P OJ t<^ 1 ft'-- - — -' ^_^ -P 1 CO 2 o^ ir\ 03 1 O -— ' • • 0) t^ 00 e 1 O >, ^— CM 1 -P 0) 1— ( 1 u 1 1 CO 1 0) ft ^^^ ■ — ^ •H 1 E + + H 1 -H ---» CTi ,— O 1 -P IS • • c o o ■H 1 X CM r^ pa 1 03 3 1 (-1 ft 1 3 .-■ — ^ 1 +J '— ^ y — N q — C 1 ft 2 (T' .,— CO 1 CO •— ' ■•' ■' '«. ^ x; 1 O O o j^ 1 >5 1 O 0) 1 f^ ^^ -p 1 -p ^J c 1 ft 1 1 03 1 0) 1 ^-^ Sh 1 e o 1 + + 03 1 -H ^~v 1 , — , — t»H 1 +^ S • • Cw i<^> CM •H 1 >< 73 * >. f^ O 1 r— ( to ITN 1 -P 1 21 • • 1 C 1 ft 1 o O 1 CO N^> O 1 ■H 1 OI be O 1 Ck^ 1 ft 1 • 1 •H c •- 1 C %^ ■H T" 0 '^ T 0 ^ T J = = 1-P XT+P XT+P X Tt (1-P ) 1+P^XT+P X ^Tt ( 1 -P ) woopw dopw where P^ = P - P . d 0 w The gross benefit from the diet (B (T)) is the benefit derived from g each prey (see eq. 4) times the number of prey eaten: G (T)P XT^ B (T) = G (T)*N = i— -2 1 . ^ ^ 1+P^ Xt+P X'^Tt (1-P ) Foraging costs are the energetic cost spent on each captured prey plus the cost spent in pursuit of missed prey. I will assume that the metabolic rate during eating (C ) equals the metabolic rate during pursuit (C ) and that both are linear functions of time: P NT = time spent on captured prey, C NT = energy spent on captured prey. M = (l-P ) XTN = number prey missed while handling, M = (1-P ) X(T„-NT-(1-P ) XTNt ) = number of prey missed while 0 0 T w p - '' empty handed , I :'-■ ' 102 T -tM +tM = time spent on missed prey, C T = energy spent on misser prey. The net foraging benefit will be the gross benefit minus the energy spent on prey minus the energy spent waiting for prey. This waiting cost is T - T -TN = T^ - NT - (1-P ) X TNt = waiting time, i m i w p w I C = metabolic rate while waiting, C^(T^ - T^ - NT) = waiting cost. Total energy spent in foraging (E ) is therefore E- = C T + C TN + C (T^-T -TN) = (C -C )(T +TN) + C T„ I em e wTm e w m w T The net benefit per unit foraging time (B (T)) is n B (T) B^(T) - E- ----- = — ^ -'- . (7) The optimal time spent per prey item (T ) is generated by taking the op partial differential of equation 7 with respect to handling time (T-L) and setting this partial differential to zero. After some manipulation, the partial differential of equation 7 becomes 103 " = V^'"'^^Vi*Wi"^ - Vi - h ' Z. = P , X + P X^t (1-P ) , 1 d 0 p w ' Z^ = ck ln(k, ) , ^ a b Z^ = C.(l+t P, X) - C.t ^X^(1-P )(1-P ) 3 a p d dp wo Rearranging k^Z, * Zj * Z^^jT The reader should instantly recognize the fact that this is a transcendental function, for vfhich no algebraic solution for T can be found. However, T can be estimated using Newton's Method of approximation (cf. Kaplan and Lewis 1971): f(x ) V1 = ^n - ------ • ^5) f (x^) where X = nth estimate of the variable (here variable is handling t ime ) , X ^, = improved estimate of variable, ■ ' n+1 f(x ) = value of function at x (here this is the right side of eq . 6 ) , f'(x ) = derivative of f(x ). n n ,1 i M 104 I iterated equation 9 ten times for each combination of ? , P and FS. ^ o w This number of iterations gave an estimate of the optimal handling time correct to within at least .001 min. The optimal handling time was derived with the coefficients used in equation 6. The encounter rate was set to the rate at which fruit flies were fed to the antlions ( X = 1/24, 1/8, 1 and 2 flies/hr). Results As shown in chapter 2, the optimal handling time is influenced by capture probability. However, when P equals P , the optimal handling time is relatively independent of the value of the capture probability. The model predicts that as P diverges from P, , the optimal '^ 0 " handling time should decrease (Fig. 3.6). This means that the relative cost of holding on to the prey increases as the probability of missing a prey while eating increases. This effect is highest when prey encounter rate is large. A reduction in handling time below the level predicted by the deterministic model is predicted if antlions normally forage on prey that are able to escape occasionally. Unfortunately, I have no information on the overall probability of escape for the prey of M^ mobilis in the wild. Based on a comparison of the mean handling time exhibited by antlions and the predicted values from equation 8, the difference in mean capture probability (Pq-P„) should be small if the antlions are foraging optimally. f» c . E 0 >> K 0) CNI K )-l • •H ao o Cm oc CO c CO x) ■ •H S 0) 4-1 4J CC 0) 3 0) e J3 •H •H HI 4-1 ^ .-H J-i •H 4J cn x: •H •H S 3 ■TJ CO >> u .^ c 0) CC CO U ^ P. O •v ^-1 6 c c- oc 0) ^ 0) g 0) CO oc *> c cu e •1-1 i-H 00 i-H •H •o ,£ ON c ? o « c s: >,o 4-1 • rH —I o S 1—1 s •H tn •H -O CO AJ CC S O.X) o 0 4J u <^ o. , OC » •H vO (U CD !-< s m 3 4J u 0) a . 0) 1-1 CC 3 3 u 6C II >^ •H Or-I fi- PM fi. 106 do .—1 0 M e a -^ z 3 a. < • ( O OE — ' C 0.6 ,. , 0.4 3 I do do 107 General Discussion Antlions exhibit two responses to changes in prey encounter rate: (1) they reduce handling time, and (2) they decrease the percent extracted i-rom each prey. The second response contradicts the predictions of the DHL model (Griffiths 1982) and the GLM model (Rolling 1965; Johnson et al. 1975; also see chapter 2). This reduction in percent extraction was predicted by an optimality model derived for antlion prey utilization. Thus, antlions appear to be making decisions that are intrinsically more complicated than the simple "eat-until-you-are-full" rule posited by Griffiths (1982) and Johnson et al. (1975). A result of these decision rules is that antlions increase ^- the net benefit derived from their diet compared to the mechanistic rules. The reduction in handling time with an increase in feeding rate is caused by three factors. Two of the factors are predicted by the deterministic model that describes the conditions of the experiment. These are: (l) the increase in extraction rate, and (2) a decrease in the percent extracted as encounter rate increases. However, the decrease in handling time was larger than expected based on just these two factors that were included in the model. One additional factor that "' may explain this discrepancy is the variance in percent extracted. There may be no energetic difference between handling the optimal amount of time (predicted by equation 6) and handling the amount of time exhibited by antlions. If this is true, then we might expect antlions - i to lower handling time if factors other than the energetics of eating - j .;'> are influenced by a change in handling time. (3) For antlions, an 1 108 increase in handling time may decrease the probability of catching the next prey. This potential loss of prey should lower handling. Thus, giyen a range of handling times over which there is no real difference in net energy derived from eating, the antlion should choose the lowest handling time. Of course, if there is a large difference between capture probability while prey handling and when empty handed, then the antlion may discard the prey much earlier than predicted by the deterministic model (Fig. 3.6). In this case, antlions should reduce the energy derived from any single prey to increase an expected total net energy derived from all prey. In two similar papers, Sih (1980), and Cook and Gockrell (1978) suggested that partial prey consumption could be studied by the use of optimal foraging theory. More specifically, they suggested that the marginal value theorem (Charnov 1976) could be used as a tool to help understand the specific behaviors exhibited by animals that partially consume prey. The insight in these studies was not that animals may be thought of as optimizers, since by now the idea is a relatively old one (cf. Schoener 1971), but that the techniques developed in other areas of foraging behavior may be useful in understanding partial prey consumption in particular. Unfortunately, the marginal value theorem is not as generally applicable as has been suggested, but this problem is easily, although somewhat tediously, corrected (see chapter 2). On the other hand, the mechanistic models proposed by several authors (chapter 2), failed in my study on antlions, even though the antlion system was the primary focus of one of these models (Griffiths 1982). I do not mean to imply that animals never use simple mechanisms, such as gut filling, as a cue to cease feeding. Predators such as damselfly larvae 109 (Johnson, et al. 1975) and Rhodnius bugs (van der Kloot I960) probably do. However, the specific mechanisms that direct foraging behavior in animals should be evolutionarily labile, since different mechanisms will work well under certain conditions and poorly under others. For example, gut filling rules for partial prey consumption may be poorly adapted under conditions where prey density is low, but prey body size is large (Cook and Cockrell 1978). It will also be poorly adapted under conditions where prey encounter rate is relatively high, but prey body size is small. Thus, animals living under different conditions should evolve 'rules-of- thumb' that will optimize prey utilization under those conditions. In other words, we should expect the behavior of an organism to evolve such that the mechanisms underlying foraging maximize the fitness of the organism. In a sense, the mechanisms should to some degree conform to the output of the behavior, since it is the behavioral phenotype on which selection acts. Animals have been consistently shown to exhibit foraging behaviors appropriate to the conditions under which they are tested. This fact suggests that natural selection is a factor in refining the prey utilization of foragers. The technique of developing optimal foraging models to study foraging behavior has been a major contribution in our ability to address this field of study. Animals have also been consistently shown not to conform precisely to the predictions from these models. Partial preference is a perfect example (Krebs et al. 1977). Handling time in antlions is a more specific example. But the optimization technique allows us to set up testable hypotheses to study further refinements of our perception of these systems. ^ CHAPTER IV THE ROLE OF FORAGING TIME CONSTRAINTS AND VARIABLE PREY ENCOUNTER IN OPTIMAL DIET CHOICE Introduction A variety of optimal foraging models has been proposed within the last decade (see Schoener 1971; Pyke et al. 1977). Many of these models have dealt specifically with diet choice and generate the same general predictions (Charnov 1976b; Pulliam 1974; Werner and Hall 1974): (1) prey types should be ranked according to the ratio E./h. (the energy derived from a prey divided by its handling time), (2) a prey type should be added to the diet solely on the basis of the absolute frequency of encounter with higher ranking prey types, and (3) prey types should either be eaten on every occasion or not at all. However, there are conditions under which these predictions change. For example, prediction 3 may be violated if the diet choice of the forager is limited by nutrient constraints (Pulliam 1975; Rapport 1971, 1980; Westoby 1974). Prediction number 2 is violated when prey recognition time is relatively long or when handling time varies through learning (Elner and Hughes 1978; Estabrook and Dunham 1976; Hughes 1979; Xrebs 1978). The most widely used foraging model is a variation of Holling's disk equation derived by Charnov ( 1976b). This model has been used to •.*j fJ study diet choice in a variety of organisms, including insects, ■>."^« ••J '■i^ 110 Ill gastropods, birds and mammals (Charnov 1976b; Dunstone and O'Connor 1979; Hughes 1979; Xrebs et al. 1977; Palmer 1981). The model is based on the premise that foragers choose prey types that maximise the net rate of return of the limiting currency (or currencies) (see Pyke et al. 1977). For this discussion, I will assume that there is selection on the predator to choose a diet which maximizes the net rate of energy intake. This expected rate of intake is the total net energy gained from foraging (E) divided by the total time spent foraging (T ): E X.E.P. E i ^ ^ ^ T^ 1 + Z A.h.P. T .111 1 (1) where E. = energy derived from prey i, X. = encounter rate of prey i during the search time, h. = handling time of prey i, P. = probability of attacking i when i is encountered. From this model, Charnov (l976b) proved that prey i should be added to the diet in order of rank until E. E * -i < -^, (2) h. T* 1 where E^*/t* is the maximum net rate of energy intake. Thus, decisions concerning the addition of prey items to the diet should be based on th« effect of these decisions on the energetic return from the entire diet. B-.:- 112 There is another way of phrasing the decision rule derived from (1). Instead of asking what prey items would increase the rate of energetic return from the diet (as in eq. 2), the effect of each foraging decision can be calculated more directly (see Cost Model helow). Although the two methods are mathematically similar, I will show that an emphasis on the costs associated with the foraging decision can lead to two further predictions about diet choice. The first is that the time available for foraging may affect diet choice. The second is that if some of the assumptions of equation (1) are relaxed, then several characteristics of the prey distribution may be exploitable by the forager to increase net benefit from the diet. I show that under these conditions, a forager's perception will affect the predictions generated from an optimization model. The Cost Model Assume that a forager feeds on two prey types, one of low quality and the other of high quality. Assume also that the forager encounters prey individually, handles them one at a time and recognizes prey instantly. When both prey types are present at high densities, there should be a high probability of missing the opportunity to capture high quality prey if the predator were to pursue and eat low quality prey. In fact, the "cost" of taking poor quality prey can be evaluated as the potential loss of high quality prey items missed while handling the poor prey. Conversely, the predator should take a poor prey item, regardless of prey density, if it could be assured that no better prey would be missed while handling this poorer prey. Any decision involves the 113 coramitrnent of a given period of time to a particular course of action (Brockmann, et al. 1979). In foraging, a decision is a costly one only if the forager misses a more profitable course of action otherwise available during that period of time. I will first derive this "cost" for a one-prey situation, then expand it to include a second prey type. Let X. be the expected harvest rate of prey i over foraging time T (when every prey encountered is eaten). I will assume that prey show a Poisson distribution with a mean of X.T (where X. is the encounter rate of prey i over T^). For a diet composed of one prey, the expected number of prey taken in time T is X.T = X .T^ - M. .X.T^, (3) 1 T 1 T 11 1 T where M = the expected number of other prey i missed while handling the captured prey i during h. (the total amount of time spent capturing and eating i, then resuming search). The expected harvest rate will be the encounter rate discounted by the rate at which prey are missed while handling captured prey. This will hold for both sit-and-wait predators and active foragers as long as prey show a Poisson distribution throughout the foraging bout. Solving for X. X = i — - . (4) 1 + M. . 11 Since M ii '^i*'^i ^°^ randomly distributed prey, then 114 E, \ , E 1 i X.S. = -i-i- 1 + A .h. T„ 1 1 1 for one prey type. Thus, equation 1 and equation 3 are equivalent even though they address the foraging decision in different ways. If more prey are added to the diet, then equation 3 expands to X. = X. - X.M. . - Z X.M. ., 11 1 11 J ij' where M^ = the number of prey i "missed" while feeding on prey j. If prey i is the highest quality prey, then in a two-prey system, M is ij the relative "cost" of adding the low quality prey (j) to the diet, in terms of the alternative decision of specializing on prey i. More specifically, "cost" is a function of the energetic return from the poorer prey item minus the expected return from the higher quality prey that was missed while handling the low quality prey. To simplify the formula, let the handling times and encounter rates of the two prey be equivalent (so that h. = h. = h and A . = A . = X ). Therefore 1 J 1 J M = M = M. Then for a two-prey diet, the expected energetic return from the diet is E A (E. + E.) — = X.E. + X.E. = ^- T ^ ^ '^ "^ 1 + 2M Under these simple conditions, the predator sliouli specialize (eat only •Va prey i) if the energetic return from a single-prey diet is higher than ' '*'^ w^r ^ 115 the return from a diet of both prey types; A,i, X(E. + E.) ^ or 1 + M . . 1 + 2M 11 IT _ V If we assume that M. . = A .h., Charnov's model can be similarly ij 1 J expanded to E.h. M. . > — -^-i! . (5) ^^ E.h. - E.h. 1 J J 1 As before, the "cost" of eating prey j is a function of the number of high quality prey missed while handling lower quality prey (M. .). Equation 5 shows that M. . is a function of the relative energetic content and handling times of the two prey types. So if the forager monitors only mean encounter rates and prey are distributed randomly, then the rule on which this cost function is based generates the same three general predictions as those listed above. But under what conditions might this cost function vary? I will focus on two such conditions. (1) If the foraging bout becomes exceedingly short due to constraints on foraging time, the number of higher quality prey otherwise available to the forager while handling a poor prey will decrease (here foraging bout is defined as the amount of uninterrupted time a forager spends searching, pursuing and eating prey). The basis of this statement will be discussed below. (2) If the forager had more 116 information about the distribution of its prey, then it may be able to judge variance in 14. .. In this case, partial prey preference is predicted. (1) Foraging Time Constraints Total foraging time for a variety of foragers is often broken into relatively short bouts. For example, a bird in an area of high predation risk may often stop foraging to scan for predators (Caraco 1979; Powell 1974). Other foragers, such as some aquatic insects, may often interrupt foraging to flee from predators (Sih 1980b). Also, the foraging of intertidal animals, such as predatory gastropods, will be disrupted by tides (Menge 1974). If these foraging bouts are confined to an interval of time t, the optimal diet may be affected by the magnitude of bout length (contrary to the predictions from equations 1 and 5), as demonstrated below. I am assuming here that the factor(s) that constrains t is extrinsic to decisions about diet choice. For the following model, I assume that the forager captures only one prey per bout. This may occur either because t is sufficiently short to allow the forager to capture and eat only one prey, or because the forager captures only one prey regardless of the length of t. For example, house wrens and many other birds foraging to feed nestlings might capture only one prey per foraging bout before returning to the nest (Bent, 1964). Here the "cost" associated with foraging on lower ranking prey is not the number of prey missed while handling these prey (as in equation 5), but rather a function of the probability of missing 117 any higher ranking prey within the time left for foraging in the bout, given that a lower ranking prey has been taken. The length of t can affect foraging in two different ways. If the foraging time available per bout must include handling time, then the time available for searching will include only the interval [0,t J, * where t is the end of the foraging bout (t) minus the handling time * for prey i (h.). No prey item should be caught in the interval [t ,tj since the predator would not have enough time to handle it. Alternatively, if the predator can extend handling time past t, such that it can attend to other requirements while handling the prey, then the time available for searching includes the entire interval [0,tj. However, no prey are available after t and the cost of foraging, in terms of handling time, will terminate at t. In either case, the time available for foraging may affect diet choice. ?or a two prey system, let g be a higher (good) quality prey and let b be a lower (bad) quality prey. Also let P(Z -t*(X^+X ) = (1 _ e ^ ) - (--i— ( 1- e ^ ^ )) . The first part of this expression is the probability that a g will * * arrive in t (P(Z \ ' ^^h^ ' h ' ^^\) • This is also trae if handling time can extend beyond t, in which case t * is substituted for t since the time available for searching is t. Rearranging the above inequality Eg • (P(Zg E^ • P(Y^) , where P(Z -— . (6) b g g g Equation 6 generates two predictions that are different than those of equation 5. (1 ) Diet choice depends on the amount of time available * for foraging (t ). As available foraging time decreases, diet breadth should increase, because the "cost" of catching low-ranking prey (i.e. the number of high-quality prey missed) decreases as t decreases, ''J independent of the encounter rate of either prey type. The "cost" is ■- c" 121 (in part) weighed against P(Y, ), which also decreases with decreasing ,.■ •' * * t . However, ?(Y . ■u ao — ( c ,—t •H •1-1 >-l ^ 0) 03 4-1 ^ c Q 3 H O o. o r* aj 3 r- Zi M-4 o c >> u •H ca ^ u n) XI QJ o j= jj 4J a C V o x: u 00 c >. •H X2 j: a -d !-i 0) CC 13 QJ •H cn > •H M •n o 00 TO N •H XI to >- > s— ' « PU >-• 0) 6 >. •H OJ 4J 1-1 c- • U-i ^— V O 60 ^-^ C Xi 4-J T-i >H O ^ ^-^ OJ 3 P- U-J n: w O-i (-1 0) >, JZ 0) QJ 00 v-i j: •H O. H X 00 ca c • •1-1 i-i 00^ >-a c CO to c tu CT! u CO I— i U 0) o c f^ 4-1 u- Q) CO 3 II ■H 0 o C r" X r" OJ 4-J !^)- 00 w o C c 1« •H 0) • n 60 x: 4-1 to CO 3 OJ 0) (-4 •H 4-1 o CO ■o CO U-I j-i M v^ dJ 4-J II •H CO !-i 00 73 •H 0) c • iH 4-1 «H 4-1 4J CO c _r^ X CO •H 3 U cu •H a o ^ 4-1 i-l OJ o CO CO a c OJ Q) u CO 0) CO 0) (U CO c g 0) S-i 0) o J-l o en oc 14-1 CO U-I 6 T) X! QJ CU c >>'< 1— < 4-1 CO M X Vj TD CO U-I u QJ C rH o CO C CO •H •H QJ CO en i-H II ao> c CO O^ CO o •H (U •H CJ a-' 0) J-i CJ COr-l e •H O- W II •H r^ M XI 4-1 •H J3 II U-I c • II •K 0) cu 4-1 fJC4-l TD • *\ OJ OJ iH ^^ 4= JS 00 II >^ 4J ^ 43 ^H Vj CO •H W OJ 3 0) C^ > U-I u CO • #. •H c u (^ 4J Vj 0) o II O O ^ "4-1 00 0) fe- , •H B , •H • u QJ 4-1 CN Q. 4J !-l • Cfl B. 00 -J- Qj •H c c t-t >^ •(-• CJ 0 CO J-l 1 — i u U •H -a 3 >> QJ ^ c 001-^ c CO CO •H c (U 3 ^ fc o 00 CT mr.r" 125 •^ ^, ^ ^-^ — CO — — cc o C<| "^ - - o - o CD 1 0 o 1 1 1 z : 1 > en <) 0 1 1 1 1 i ^ o «i *— UJ UJ I— Q. O UJ WO O GO — T" 6 d — r- CVJ 6 N303 - o 03dS- — r- CM d J/1 ^ CO tn i-H •T-t c to e o 4-1 •3 S-i 3 cc 0) 01 "H •H C o tn g k* U-f 3 CO P- >-l ^v 0) -H 3 s: 13 o a (u u^ c 1— t U-I (1) x: u-i 00 OJ 4-1 • 3 C4J -H -i-l e o i-H O . .•— s m -D tn •H c II J= vXI W tn 4-1 XI tn ■ >-• ■— -H >, -u T3 a .-H O o 1—1 4-1 c JJ 4J U C i-H 3 ., 0) o CO c c O ^ I.M OJ S-1 CT a ^ -H w 00 .^ QJ 0) aj cj c CO c CO 1— t x: a CO M^ 1-1 u II 3 >^ 05 W -H J-i 3 o -Q X tn 4J 14.4 JZ 60 1 to tn 0) 11 ^ •3 . .-^ (U ^ 01 60 0) C '-N CO x: s: Q) 4J OJ ij (U c o H r- a dJ -w' MO) *H 4J •H 3 a sj 4-1 • t: 4-) tn 1 (U 01 • XI tn 01 a ^ OU-4 CO ^— s ^ 3 M ^4-1 CU Cw u 60 o & M D--H OJ C II x: c c tn-TD c O tn tn (-1 (U w a; 1—1 60 •H 3 ,-1 ^^ ;j r< ^^ 4-J C 1—1 tn LO /'-^ 0) j_i tn (-1 3 oj tn ij •H O ^-^ O CJ 0) o c -3 U J2 C ^-1 4-1 3 tn Q) -3 •H •u 0 3 (DO CO 4-> •H ^ o 14-1 4J 3 c QJ 3 ^- y> C C cr 01 u CO ^ •H 4-1 01 tn en QJ — to a tn OJ 3 O QJ tn J (-1 0) &. tn C H X o 0) a en QJ i-H CO H C iJ-t 0) c !-i CO M 0) 14-1 Vj 1—1 •H • )-l -H o i-H S-i 0) 4-1 • 01 "a tn CO 4-> )-i X /-^ U-l 0) •-H a) 01 in Uh S-i V-l • >-4 ^ 3 ^ ■ — •' •H o to ^-v 4-1 o -O in 3 -u •H tn Q) c o" en x: ^^ U 0) o TJ 4-1 tn 01 o 3 OJ tn •H 0) 0) -iJ CO 4-1 4-1 4-1 -H c 1 M c c tn to CO T3 01 U •H 3 e 3 iH D.~~^ 0 s- cr 3 ^ O • in O 0) QJ 6 C "-1 O • C 4J •H 0) O ^^ u >> tn 60 • • e II 1+-4 X M tn c; o 4-1 0 ^-^ i-H 11 u • ■3 CO D.14-1 u 4j tn QJ ro 4J -H o c 4J • tn >-4 4-1 4J U-I 4-1 O tj -3- -H 4-1 CO c tn -H •H t—l 0) 4.1 CO 4J •3 • 0) CO C C (J o. a. -ri OJ in i- S-4 u-i OJ QJ c u . 3 0) rH E 14-1 O >> -H o-o 60 c c U-l >< Q) U-I II ■H 01 «-! Vj •H 0) >-i 01 tn -K fn 00 O i^^ -a ^ — C1.T3 CO 4-1 128 1— 3 O IS - o X o ^ '^ in < ts — CVJ ^ - (T) 1 1 (T) >- UJ UJ »- Q. O UJ l/l Q DD < q: or UJ Ltl CO K r UJ I- UJ O UJ O Q OD — I— — 1— — r- C\J o» cv. CO o o o CNJ o >- UJ q: Q. >- -I < O X UJ I- < cr h- Z) o o z UJ CJ N30g _ 03dST *1 129 some degree of partial preference to be exhibited. In other words, a prey type that is added to the diet will be attacked less than 100 percent of the tims when it is encountered. In fact, partial preference is predicted from the models derived above. For foraging bouts during which only one prey is taken, equation 6 predicts that diet choice should change as the time remaining in the foraging bout approaches zero. Over many bouts, a forager will exhibit partial preference due to this change in the optimal diet. In the following discussion, I consider bouts in which more than one prey is taken. Two important levels of variance may exist in M. .. The first is the variance in expected encounter rates of different prey types. The second is the variance in the actual arrival times of individual prey, in essence, the variation within the expected encounter rates. I will illustrate the importance of these sources of variation to diet choice by using two simple examples. First let two prey types be distributed in runs of three each where the time between encounters of each prey is equal. In this case, a forager might see the following prey in order of occurrence: g^g^g^b^b^b^g^g^ggb^b^bg , at times rnrnmrnrriryifyimmm m m ^r2^5 4'5'6^7 8^9-10^11^12 ' where T - T = 1. Let the benefit from prey g (S ) be 3 benefit XIX g units and E = 1. The handling times for both prey are equal: h = 130 h^ = 1 . So if the predator captures one prey it will miss the next one. If the forager eats only g, it can start with g^ , then eat g , then eat g^, etc. In the first 6 time units (T -T,) it would collect 6 benefit units. If it started with g^, it would collect 3 benefit units in the first 5 time units, but thereafter it would collect 6 benefit units for each set of 6 time units (starting with g ). So as the foraging bout gets infinitely long, the forager will get approximately 6 benefit units in 6 time units. If the forager eats both g and b, it can start with g^ , then eat g , b^, g,, and so on. With this sequence it gathers 7 benefit units in 6 time units. It can start with g^, then eat b^ , b , g , etc. This sequence produces 5 benefit units in 6 time units. So on average, the predator will take in (7/6 + 5/6)/2=1 benefit unit per unit time. Now suppose the forager knows the temporal placement of several prey simultaneously. It should eat any g that it encounters since it can do no better, but it should capture b only if no better alternative is available. If the b is followed by another b, it should eat the first prey, because the time spent waiting for the second prey (assuming the predator decides not to take the first prey) should be incorporated into the handling time of the second prey. Here S/h for the first prey is 1 and E/(h + 1 ) for the second prey is 0.5. On the other hand, if the second prey is a g, then E/(h + 1 ) for the second prey is 3/2 so the forager should skip the first prey (b) and eat the second prey (g). This means that the relative values of b , b , and b^ are different (actually b = b / b_). Thus, if the predator could monitor the relative cost of taking each prey, it would take either of the first two of a prey b sequence but never the last. This diet would yield 7 benefit units in 6 time !•> 131 units. Thus, if the animal's foraging decisions were consistent with the Charnor equation then it should on average take in 1 benefit unit per time unit. Under the simple conditions of the example above, if the predator can monitor several prey at a time, it should always do at least as well as Charnov's rule (eq. 2) and may do better when using a decision rule that takes into account the relative temporal position of prey. If there is any variation in the encounter rate of prey, the effect of averaging on diet choice (as illustrated above) will be important even if the predator can not simultaneously detect a number of prey. This can be demonstrated by simulating the relative effect of the time the animal estimates M. . on the net energetic intake when the mean encounter rate of prey varies. I simulated three different decisions for a two-prey system: (1) take both prey, (2) take only the higher quality prey, and (3) take lower quality prey when E,h, < M. . < ^-i. E.h . - E .h. 1 J J 1 I varied the encounter rate of high ranking prey (X ) as a sine cr o function with a range of zero to x (where x varied from 0 to 3.0) (see Fig. 4.4; note that the abscissa is the mean value of the range of X ). The simulations illustrate two important properties about decision rules: (l) the forager always does better if it instantaneously monitors M. . than if it follows a rule in which it bases diet choice on the average encounter rate in the foraging bout, and (2) if the encounter rate of high ranking prey varies, we may expect ■^" 13 4-1 r-t tn o O •H o 3 )-< OJ r-i. j: rH CN o 0) a cn CO I-- ^J tn 4-) 4-1 •H >> tu •H s-l CO QJ c CO 4-1 1-1 u O 13 CO Sw' •H 3 4^ 0) r" o; 0) QJ 'O O •H O > 4-> a. 4-1 QJ 4J X • o C vO cu cn tn f— tn H 00 CO c 3 rn 0) CO H •H C OJ 0) ^ £ E rH rH •H r-l 0) S-i OJ H o n • E 3 c e oj > M c ■ •iH y^"\ 3 U CO •H > 0) M-l o tn ■^ 1 oocn 0) 4-> O .-H • •H rH Q)t-< 'J5 c 6 O ^ 4J CO C ■ — •' CO o O 003 !0 •H tn •H to 1—1 r< a; r-t U T-^ M tu 4-1 a ^ II —^ ^ 4-1 o r-A QJ •H ^^ >> c X3 s 4-1 a •-H U 4-1 (-1 CO CO r< tn •rt 13 >, ■^ 0) •H E • ^ 13 4-) tn X OJ •u l-^ r^ X CO •H rH a 00 H c X. u • 5 c >, O OJ OJ 4-1 o 00 o rH l4-< CO oc • #1 00 X rH rH CO oc . rH CO QJ E E CO 0) 3 ^ -^ 0) II i- 4.i U_l c o 2 (U 00 !- X o O ^ s-l 4-1 >- >^ r< OJ w U-l QJ <4-l (4-1 O x: dl • — ' 3 QJ 3 CO 4-1 >- 4-i • • A Q) tn to X 00 a C ^ o(ro x: Si o r< -o E CO r-^ << II 4-1 *> E 4J 1—1 o ?-. ■u + OO tn •H r^ 3 i-l 4J tn --V 1^ w o E Q) 3 CO O y-i •H c .-^ c cn u X tn 0) X iH O to ■ • QJ 4-1 E tn (U CO o > tn T3 in 4-1 T3 -i 3 0 C O 4-1 00 3 C s 0) c i.-< -H ^ 0) E •H 4-1 o X u QJ 4J •u Uj 4-i CO 4-1 C 14-4 QJ o C •W •H <1) V4 QJ •H /^^ U U-J OJ iH • • TD E CO 13 o QJ c QJ to /-> X CO a a X o ,^ QJ 4J 3 >^ to J-i !-i 4-1 •H 3 X 0) cr 1-H E vO CO 01 O X 4-) 4-1 c CJ -^ m a. tn PH u OJ to 4-1 r- > c QJ CO u 3 c 00 l« "ob •H -H ri 00 x: QJ QJ cr •H c o •H 4J cn cn c 4-1 ■ X QJ O •rl j: o --^ 4J •H U~l T3 3 o. u en fU •H 3 U . c > 3 c J= D. •• C o 0 C to 4-1 o QJ •o o 4J tn c 3 rH U-! 11 c c x. •H O cu o r-t 00^ •H t-l 4J Q) U jQ U -H OJ o ,^-^ <-c tn 0 CO OO CO v_/ 4-1 E U-, LO QJ a-x tn r; 1-1 (-J CO •H C ^—^ CJ •H 5 3 O XI 3 4-1 0) C 0) tj Q) x; e U-l .-< cr ^ r^ o X p^ ^ QJ yr*^ o •H CJ o E- •»-( 3 y 4-t X < M Vj X) ^c 4J 4-1 '^ ' 4-1 0) C QJ m CO 4J CN cn O 4-) CO JZ • 3 tn \»^ •H >^T3 c • c 4-1 M ^-^ o- •H X CU *-• QJ QJ •H 3 O 3 >, tU r-H 4.1 0) 00 0) ^ "O o •tH 60 JQ o» o ^i UJ >. N G 2~ < z -) < lALI ON O is UJ , , < liJ l_ o y- -I UJ a: z < UJ - 1 1 ft O h 1^ o I 00 6 CO d d CVJ d o in CVJ O O o d u> (3l^ll/llJ3N3a) 3>^V1NI A9a3N3 dO 31Vd 13N 134 to see the diet change through a foraging bout. In other words, the predator may exhibit "runs" of differing diet choice. These simulations also illustrate that the amount of time the forager takes to estimate X will affect the energetic intake from the diet. For example, if the forager simulated in Fig. 4.4 was to estimate A based on one cycle (360 time units), then it would base diet choice decisions on the mean encounter rate from the cycle. For a range of 0 to 2, the mean would be 1.0. Using the mean encounter rate, Charnov's equation (2) suggests that the simulated animal should switch from generalist to specialist at a mean X of 0.5. Yet the g simulations show that this decision underestimates the best switching point. In fact, the switch from generalist to specialist should be closer to a mean X of 1.50 (see Fig. 4.4). The basis of this underestimation can be viewed in terms of M. .. With X constant, most b prey will be taken when the arrival rate of g is low and when '''^ij ^^ "'"^"" ''^^^ ^'^^ arrival rate of g is high, the probability of taking b will be small. Thus, summing over all prey b taken, the number of higher ranking prey missed while handling the lower ranking prey will be smaller than expected when the arrival rate of g is constant. In other words, M is reduced when variance in prey encounter rate increases. This reduction shifts the optimal switching point to higher levels of X . O A second factor is important when the mean encounter rate varies; the calculation of E/T from equation 2 is overestimated for both the generalist and specialist diets. If X is varied as a sine function g (as above), then the amplitude of the sine wave can be used to simulate variance in prey encounter rate. As the amplitude increases, the net 135 energy derived from any of the simulated diets decreases (Fig. 4-5). This is because the energetic intake rate (E/T^) as a function of X g changes more rapidly at low values of X than at high values. When the mean value of a is used to calculate E/T , this nonlinearlity will cause E/T to be overestimated. This can be shown as follows: let E , E , h , and h, be constants. Also let A be a constant g b g b b (since the level of X will not affect diet choice anyway under the assumptions of the model). The energetic intake for a two-prey system is E EX + E^ X ^ ^ g__g b b T 1 + h X + h X T ""g g \ b The change in E/T with changing levels of X is 9 f El E (1 + h^ X ^) - E^h -X ^ ___ — = -i -— - Li__^ . (7) 9A I'T^ (1 + h^A ^ + h A )^ g I T J b b g g' This function decreases monotonically with increasing A (Fig. 4.6). If Charnov's rule (eq. 2) is considered an estimate of E/T^, then i under most circumstances the forager will overestimate the net energetic return from both the generalist and specialist diets if it requires a relatively long time to estimate A . Thus, the time the forager takes o to respond to changes in prey densities should vary inversely with the rate of change in mean prey density. Since a number of predators have been shown to respond to changes in prey density (Elner and Hughes 1978; Giller 198Q; Goss-Custard 1981; Jaeger and Barnard 1581; Krebs et al. 1977; Sih 1980a; Werner and Hall 1974), it is not unreasonable to Figure 4.5. Simulations of net energetic intake from three foraging decision rules when the variance in encounter rate of high ranking prey- changes. Prey encounter for both high quality and low quality prey was random with a mean encounter rate of Xg and a^^ (respectively) . X^ varied every 10 time units following the equation: (r- sin (max. time in interval)+l) • (mean g over 360 time units). Thus, r is a function of the variance in Ag. Mean X =1.0 for all simulations. The threshold Mj^^ (see equation 5) for these parameters is 0.5. All other parameters as in Fig. 4. (i«r'-T« ' 137 UJ SYMBOL X X 0—0 FORAGING RULE GENERALIZE (EAT g AND b) SPECIALIZE (EAT ONLY g) FOLLOW M ij EQUATION 5 RULE (SEE IN TEXT) 0.2 0.4 0.6 0.8 1.0 VARIATION OF ^g (r) 0) oo IJ c CO •H 60 00 C cn c •T-( 0) CO 3 4-1 J= o CO o r-4 Pi X o 4-1 y-l • •H 4-1 3 OJ X ,^ OJ ^-s J^ 4-1 iH o; r" OJ -o o 0) o tn e -a tu cn tn en E »■ CO u > ,fl di o 4J e OJ ^1 S-i U-i n CO O x: u en cn c c E o o o •H •H ^J 4J 4J U-l O •H s_^ •H c TJ •H . XI 60 >. X !-l OJ II OJ ^1 00 C D.^ a; ^w^ 4-1 >. u-1 • to aj >, vO ^4 ■-H X: ^U ^-1 ^4 T3 - ~l ^ _l < _l O < 2 f- UJ 2 < Ixl Q- o> LU o» Q CO ^-^^ O —' _l o Q ^ OQ 1 S 1 > o >l: o q CM o> << > UJ GC Q. O < DC X o o LU < QC CC LU o o Z LU e mfi 140 suspect that at least some foragers may be able to monitor M . This should hold true especially for predators that hunt by sight for prey which arrive simultaneously (such as planktivorous fish or insects, salamanders, birds foraging in swarms of insects, etc.)- Discussion The models presented here generate the following predictions: (l) if foraging is confined to a given period of time, t, then as t decreases, diet choice should become less specialized. (2) Diet choice should reflect variance in prey distributions if this variance affects the variation in the relative "cost" of capturing and eating low ranking prey. This cost in a two-prey system, M. ., is the number of higher quality prey missed while handling lower quality prey. Short term changes in prey encounter rates should be incorporated in the decisions the animal makes about diet choice. This is because decisions based on long term averages will yield a lower return than decisions based on short term averages. Thus, variance in the number of prey added to the diet should increase as variance in prey distribution rises. As a result of the variance in diet composition, partial preference should also increase with an increase in the variance in prey distribution. (3) Diet choice will also be strongly affected by the number of prey the predator can monitor at any one time. If several prey can be perceived at once, then the forager should be able to estimate M. . nearly instantaneously. In this case, the optimal diet changes rapidly and as a result, partial preference should be exhibited. Predictions (2) and (3) above illustrate that how the forager 141 estimates prey distributions will strongly affect the optimal diet. Clumping of prey should select for foraging strategies that restrict the time required to estimate local prey density (e.g. through short term rules of thumb as suggested for granivorous birds by Barnard 1980). Common prey should allow predators to make foraging decisions based on accurate estimates of M. . if more than one prey can be seen during the time of the prey choice decision. The importance of these predictions is illustrated especially in studies on switching in which a predator tends to forage disproportionately on prey that are most abundant (Murdoch 1969; Murdoch et al. 1975). Several authors have shown that switching should occur when some learned foraging response is exhibited by a forager (Hughes 1979; Murdoch 1969; Oaten and Murdoch 1975). But in some cases, switching may be predicted if the forager can quickly monitor changes in prey density, irrespective of any learned behavior. Visual predators such as fish (Bernstein and Jung 1979; Olmstead et al. 1979) and salamanders (Jaeger and Barnard 1981) have been shown to respond quickly to changes in prey density and may therefore be able to monitor M closely (equation 5). One prediction that stems from these models is that predators should either use visual estimates of prey density (as shown for salamanders by Jaeger and Barnard 1981), or should restrict the amount of time over which they estimate prey encounter rate if visual estimates are impossible. Thus, the length of the forager's "memory window" (sensu Cowie 1977) should be set small enough to minimize the effect of averaging unless the costs of misjudging encounter rates outweighs the costs of averaging. If foragers are estimating M. . based on short term rules, then as variance in prey encounter rate increases, the number of prey included 4 1 A ■-4 142 in the diet will vary. As a result, foragers should exhibit a higher degree of partial preference (see Pulliam 1974). Clearly the foraging decisions exhibited by animals may be much more subtle than foraging theory has allowed. To understand exactly the nature of diet choice decisions, we need much more detailed data on foraging behavior. Three studies support the first prediction. Jaeger et al. (1981) analyzed feeding selectivity of salamanders foraging on large and small flies. They showed that a salamander on its own territory preferentially fed on large flies, but on unmarked territories or on a conspecific's territory, the salamander exhibited no preference for either fly type. They also showed that a territory owner spends most of its time foraging. A salamander on no territory or one on a conspecific's territory spends a large amount of time either marking or displaying submissive postures, and feeding is interspersed between these two behaviors (R.G. Jaeger, pers. comm.). The conclusion Jaeger et al. (198I; pg. 1100) reached was that salamanders sacrificed "initial caloric yield until they had established marked territories and then (switched) to a higher sustained caloric yield". Jaeger et al. (I98I) viewed the foraging bout as the entire time spent in an area. However, the decision of diet choice should not be based on time spent in other activities but solely on time spent foraging. When analyzed in this way, on average only one fly was taken between bouts of marking (when the predator was foraging equally on large and small flies), not including the time spent in a submissive posture. When the predator was foraging selectively, the foraging bout was uninterrupted. The salamander data can be reinterpreted as follows: the length of the ,jj foraging bout was long when the salamander foraged on its own territory ' ^ 143 but extremely short when on another territory or on an unmarked territory. In the latter two cases, foraging bouts were broken up by marking or displaying activities. Thus, if the salamander were to maximize its net rate of energy return while foraging, we would predict -' ; that it should increase its degree of specialization with an increase in foraging bout length. "*» Freed (1981 ) analyzed diet choice by house wrens foraging for nestlings. When in the presence of a nestling predator (the fox snake), . ' ' the wrens decreased the amount of time they spent foraging and fed their nestlings smaller prey than when no predator was present in the area. '\^- Thus, as foraging bout time decreased, diet selectivity was also . ' ;'' reduced. Freed (1981 ) suggested that, based on these data, the decision rules under which the wrens were choosing their prey changed when a : r predator entered the area. However, if the rule governing diet choice is to minimize the "cost" of foraging on low ranking prey and thus ' ■'■ maximize the net rate at which energy is taken during the foraging bout, then the models presented here suggest that the decision rules with or '• without predators are identical. The change in the wren's foraging behavior is consistent with the predictions from optimal foraging theory generated from the models presented in this paper. An exceptionally tractable system for evaluating the effects of foraging time on diet choice is found in the intertidal predators whose .'. foraging bouts are constrained by the tidal period. For example, the predatory intertidal snail, Acanthina, searches only during low tides (Menge 1974). The models I presented above are particularly suited for analyzing this species since only one prey is taken per searching bout. Menge (1974) showed that these snails exhibited strong selection for 4 LH^ specific prey early in the tidal phase. As the tide rose and the available searching time decreased, selectivity decreased (less preferred prey were taken when encountered). Here again, the optimal diet changes as foraging time changes . Thus, the relative value of any given prey type also changes as a function of foraging time. It is often useful to think of foraging behavior (as illustrated by the data from Freed, 1981, Jaeger et al. 1981 and Menge 1974) as hierarchically organized (Dawkins 1976; also see Mesarovic et al. 1970, Hassell and Southwood 1978, and Gass and Montgomerie 1981). In other words, the total set of behaviors that an animal exhibits can be viewed as organized into levels. The decisions made in each level are constrained by information from higher levels. We can dissect foraging behaviors and explicitly model foraging only if we confine the models to work within the constraints imposed by higher order effects. The strength of this organization lies in the fact that foraging models, when viewed as specific levels in a hierarchical design, become more robust than non-hierarchical foraging models. Equation (6) specifies the optimal diet when the foraging bout length is set at some time interval [0,t]. This interval may be interpreted as the boundary of the foraging model. Decisions regarding diet choice are based on factors that exist only within these boundaries. At a broader level, the length of the foraging bout is affected by constraints on the total time available for foraging (t). In the above examples, the constraints are predation pressure, intraspecif ic interactions and environmental effects. Many other studies have pointed to similar factors affecting foraging behavior and foraging time (Baker et al. 1981; Barnard 1980; Garaco 1979, 1980; Hervey 1969; Milinski and Heller 1978; Norberg 145 1981). The effects of these constraints can themselves be modeled; the output of such time-constraint models would define the boundaries of the foraging models. The foraging model (viewed in terms of one level in a hierarchy) need not be supplanted by another set of models that specifically describe a new set of decision rules for the forager when other factors impinge on foraging. Rather, this hierarchical organization of models generates predictions that should describe foraging behavior under a wide range of conditions and, therefore, may be of more heuristic value than specific models based on unique conditions. Summary Predictions generated from optimality models are inescapably based on a number of assumptions. The predictive value of these models is often determined by the degree to which the behavior of an organism fits the underlying assumptions of the model. I analyzed optimal diet choice by relaxing two sets of assumptions made in previous optimality models. (1 ) Foraging-bout length fthe uninterrupted time devoted just to foraging), generally treated as infinitely long, was shown to affect optimal diet choice. For many foragers, foraging-bout length may be considerably shortened by the presence of predators, or by physical or social features of the forager's environment. A model was derived which incorporates a short bout length into the decision of diet choice. The model predicts that animals should become more catholic in their diet choice as the amount of uninterrupted foraging time decreases. This prediction appears to be supported by three studies from the literature. 1^-' 146 Jaeger et al. (1981) showed that salamanders incorporated more lower ranked prey (small flies) when they were either on the territory of a conspecific or on no territory as compared with prey choice when they were on their own territory. In this case, foraging time was uninterrupted when the salamanders were feeding selectively, but continuously interrupted by submissive behavior and marking behavior when no diet choice was exhibited. Freed (1981) showed that wrens foraging for nestlings spent less time per foraging bout when a predator was in the nesting area than when no predator was in sight. The reduction in foraging bout time correlated with a reduction in prey size fed to the young. The foraging time of some intertidal snails was shown to be confined by the length of the low tide cycle (Menge 1974). As the end of the low tide drew near, the snails decreased diet selectivity. Thus, as the remaining time available for foraging decreased, the predator exhibited a lower degree of prey selection. (2) Variance in prey encounter interval was shown to affect the utility of classical optimal diet models in predicting the optimal diet. Charnov's ( 1976b) model is shown to over-estimate the net rate of energy intake when mean encounter rate varies about some fixed level. Predictions from Charnov's model are incorrect over some ranges of prey encounter rates due to this over-estimation. I show that as variance in prey encounter rate increases, the time over which the forager estimates prey encounter rate will have a strong effect on the ability of the forager to maximize the net rate of energy intake. Foragers that forage on patchily distributed prey should use a shorter amount of time to estimate prey density than foragers that prey on evenly dispersed prey. Thus, animals that are capable of reducing the time required to estimate prey density 147 (for example, visually hunting predators in areas of high prey density) should alter their diet in response to local variation in prey density. For this type of forager, as variance in prey encounter rate increases, fluctuations in the number of prey types in the diet will increase. As a result, there should be an increase in the degree of partial prey preference exhibited by the forager with increasing variance in prey encounter rate. CHAPTER V OPTIMALITY, HIERARCHIES, AND FORAGING Introduction The conditions under which animals forage are generally complex, even in the laboratory. The animal's perception of its environment, its internal state (sensu Sibly and McFarland 1976), and numerous environmental factors may influence their behavior. This complexity has made some authors doubt the validity of optiraality models as a tool in the study of behavior (Simon 1956; Schluter 1981; Zach and Smith 1981). The issue is an important one, since if this complaint is •• . , correct, then the approach should be abandoned in favor of a technique (or techniques) that will be more useful in predicting the behavior of :-^' animals in nature. At one level the approach is clearly useful, and it has been used successfully in each of the four previous chapters of this dissertation. However, none of the existing optimality models ,^ :; adequately address the complexity of foraging behavior. I think there are two aspects to this problem. First, there is the question of generality: how robust are the predictions generated from optimality ''"■•:■'■ models? Second there is the problem of measurability: if foraging '■ decisions are based on environmental conditions, is it possible to ») 'i measure these conditions adequately? I argue that hierarchical modeling can be used to solve both problems: (l) hierarchies are useful in 148 ;.-^.-' 149 deriving generalizations about foraging decisions (or any other decisions), and (2) the use of hierarchical models may aid in decreasing the number of variables that need to be measured, and thereby reduce the problem of measurability. I begin this chapter by briefly discussing the major questions that are addressed when using optimality models, and then briefly describe characteristics of hierarchical design. After defining the question and the technique, I can then discuss generality and measurability in foraging studies. The first four chapters of this dissertation will be viewed in light of this discussion. Optimality In the introduction I suggested that optimality is a tool that can be used to address specific questions concerning behavior. However, the question is not whether animals are foraging optimally, but whether our perception of the salient features that influence foraging is sufficiently broad to predict an animal's behavior. Maynard Smith (1978) discussed this issue in detail. We can use optimality theory in three ways: (1 ) to formulate our perception, (2) to generate testable predictions from this formula, and (3) to re-evaluate our perception and add to it where it appears to be inadequate. Step three yields new, testable hypotheses. Where does this fit in the study of behavior? Tinbergen (1969) has suggested that the complete study of behavior involves four major topics. The first two topics include an analysis of the proximal mechanisms in behavior: (l ) the causation or control of behavior, and (2) the development of behavior in the individual. The y 'iit 150 se:oad two topics deal with ultimate factors in behavior: (3) the adaptive significance of behavior in relation to an animal's environment, and (4) the evolution or phylogenetic origins of behavior. Optimality theory is particularly appropriate to the study of the third topic, the adaptiveness of behavior. Although causal factors are important, I argued in Chapter III that the study of optimal behavior patterns should be made independently of the causal mechanisms, since selection should act at the level of the phenotype (the behavior) and not on its component parts. With optimality we are interested in how the animal adaptively responds to its environment, and the relationship between these adaptations and behavior. Natural selection is assumed. This assumption allows us to compare the response of an animal in a given situation with an expected response based on the selective pressures that we feel are important. A poor fit to predictions suggests that we are either incorrect or incomplete in our evaluation of the selective pressures. The mechanics of using optimality theory involve the construction and testing of mathematical or graphical models. The model is constructed around a maximization parameter, or currency (Pyke et al. 1977). The use of a given currency assumes that maximizing the net accrual rate of that currency will maximize fitness. Some models have been constructed that address the minimization of a cost function (of. Sibly and McFarland 1976). However, minimization of net cost can be considered identical to a maximization of net benefit (McCleery 1978). Although most foraging models have used energy as a currency (e.g. Krebs et al. 1977; Goss-Custard 1981; Hughes 1979; Pyke 1978), some studies have shown that energy may not be appropriate for some animals 151 (Rapport 1981; Westoby 1978; Greenstone 1979). These latter studies also indicate that more than one currency may be necessary to describe the maximization parameters (also see McCleery 1978; Sibly and McFarland 1976). In this dissertation and in all other studies of optimal foraging behavior, the measurement of the maximization parameter is determined in absolute terms. For example, a joule of energy expended in prey capture is considered to be equivalent to a joule spent traveling to a resource patch or spent during mating behavior. Odum (1982) has suggested that the importance of any absolute unit of currency will correlate with the hierarchical position of the subsystem in which the currency is spent (see "Hierarchy" below). For example, in the ecosystem one joule of puma has a higher "embodied" energy than a joule of grass. This is because the number of joules entering the system (in this case from solar input) required to generate a joule of puma is vastly larger than that required to make a joule of grass. In behavioral systems, a joule spent in early spring searching for mates after a winter hibernation may represent a larger investment than a joule spent searching for food the previous fall. This will be true if the ratio of energy expenditure (or the cost of the behavior) to fitness associated with that expenditure is lower when searching for mates than it is when searching for food. Due to relative differences in currency between subsystems, Odum (1982) suggests that energy should be expressed in terms of embodied energy and not in absolute terms. The use of embodied energy may be critical in comparing costs and benefits of hierarchically distant subsystems, but is less important when the analysis deals with any single level as is the case with most optimality models. This issue is discussed further below (see "^laximum Power and Foraging Hierarchies"). -..,. 152 If several maximization parameters are required to describe behavior, two alternatives are available. (l) All parameters can be combined into a single model of behavior. This requires that the costs and benefits associated with all parameters be measured in the same units. ?or example, the male smooth newt, Triturus vulgarly, performs a long and elaborate courtship on the bottom of a stream (McFarland 1977; Houston et al. 1977). During parts of this courtship sequence he builds up an oxygen debt and must surface to breath. The male must correctly time his courtship and breathing in order to pass his three spermatophores successfully to the female. There are three different behavioral components in the courtship sequence, two of which include different segments of the display, and the third is the rate of spermatophore transfer. These behavioral components were combined into a single model with several parameters that predict oxygen debt at any given time during the courtship. The model was based on the assumption that the newt should maximize the probability of fertilization. Thus each parameter was couched in terms of its effect on the fertilization probability. (2) Instead of a single model, each parameter or set of similar parameters can be treated separately in a series of nested models. Each submodel represents a separate behavioral decision. This is the hierarchical approach. The single model approach can be thought of as a subset of the hierarchical approach, since only one behavior or decision is addressed. In the newt example, the regulation of courtship activity as a whole could be examined with a higher level model that regulated the category of behavior that was exhibited by the animal, such as mating, feeding, or inactivity. One advantage of the 153 hierarchical approach is that each submodel can use a different currency. For example, the influence of predator pressure can be modeled through the use of time budgets. Diet choice can be modeled in. terms of calories derived from foraging during the time allotted for foraging behavior. Thus the response to predator pressure can be examined in terms of time, while diet choice can be couched in terms of energy (or nutrients). I used this approach in Chapter IV (also see Lucas 1983), but I only modeled a single hierarchical level (diet choice) by treating decisions concerning time budgets as a simple constraint on diet choice. Since hierarchies have been used for a wide variety of research fields, including ethology (Dawkins 1976; Tinbergen 1969), ecosystem dynamics (Odum 1983), evolution (Gould 1982; Lewontin 1970), foraging patch utilization (Charnov and Orians 1973; Gass and Montgomerie 1981; Hassel and Southwood 1978), and the expression of sex ratio patterns (Frank 1983), it is important to explain briefly the type of hierarchical design I am using. Hierarchy Mesarovic et al. (1970) treated hierarchical systems as a series of input/output subsystems, each at a given level of organization. The input can be thought of as cause, the output as effect. In an optimality model, the inputs are the variables, the subsystem is the model (expressed in the appropriate currency), and the output would be the behavior that resulted from the solution to the model. Mesarovic et ;^"' 154 al. (1970) suggested that for the hierarchical design to be useful, the functioning on any level must be as independent as possible of the functioning on other levels. Thus each l-,vel must be represented by a unique model, although levels will be connected by input constraints or feedback loops (Fig. 5.1 )• Each subsystem in some way controls or intervenes in the input of the system below it and may receive feedback information from that lower system, although there may be no feedback loops in some subsystems. In foraging, subsystems can use either a maximization parameter or a constraint parameter. Maximization parameters in the wren foraging example include such factors as energy, or the risk of predation. The type of subsystem that represents maximization parameters can be thought of as a specific level of behavior. More than one maximization parameter may be included in a single model, as in the case of energy and nutrient co-requirements from the diet (e.g. Pulliam 1975; Rapport 1971). Constraint parameters include factors that will modify or limit the expression of the behavior associated with lower-level subsystems. Thus, the constraint subsystems will always lie above a subsystem in which maximization parameters are evaluated. Examples of constraint parameters include tidal cycles or daylight cycles that restrict foraging time, and weather patterns that may influence migration or social behavior. Including constraint parameters in subsystems changes the original definition of the hierarchy proposed by Mesarovic et al. (1970), since these subsystems are not input/output systems in the context of behavior. However, this expansion of definition will give a model more flexibility, as I show below. ^'^' ^ 155 SYSTEM input ^ LEVEL N SUBSYSTEM intervention input input input ^ LEVEL N-1 SUBSYSTEM intervention performance feedback ^ _^ LEVEL 2 SUBSYSTEM intervention \/ performance feedback ■^ LEVEL 1 SUBSYSTEM V output v" output / output / cutout Figure 5.1. A simple hierarchical system (after Mesarovic et al. 1970) '•■•^1 156 As stated above, adjacent subsystems may or may not include f ee Iback loops. A feedback loop involves the processing of information from lower levels in the formulation of decisions at higher levels. These higher-level decisions will then constrain the lower-level processes. Thus, constraints on lower-level processes will be affected, in part, by the result of previous constraints on those levels. For simplicity, I will call a feedback subsystem a subsystem that receives information feedback from a level below it (as in level 2 in Fig. 5.1), and a non-feedback subsystem is a subsystem that receives no information from the level below it (as in level 1 in Fig 5.1). By this definition, the bottom layer of any hierarchical system will always be a non-feedback subsystem, since it does not receive information from a system below it. Maximum Power and Foraging Hierarchies The optimization principles of behavioral systems proposed here assume that the evolution of decision rules will proceed such that animals will maximize the net accrual of limiting resources or currencies. By maximizing the net harvest rate of these resources, an animal should be able to optimize investment in activities that maximize fitness. The decomposability of systems into a hierarchically nested set of subsystems, each of which can be treated as an optimality model, is implicit in the use of hierarchical design in this dissertation. Thus optimization principles apply to the whole as well as to each of the parts. The optimization of time budgeting implies an optimization of behavior within each behavioral bout. The optimization of diet 157 choice implies the optimization of prey utilization. The optimization of patch choice implies the optimization of searching behavior once in the patch. Higher level decisions impose constraints on lower level decisions, and these constraints are to some extent conditional on the behavior exhibited at lower levels. Information ("performance feedback" in Fig. 5.1) is passed from lower levels to higher levels through the behavioral output of the lower levels. The constraints imposed on lower levels can be thought of as feedback from this information flow. Another type of systems approach has been used in the analysis of ecological and economic systems. Odum (1982, citing Lotka) proposed that the design of systems should follow the maximum power principle. In other words, systems should develop designs that maximize the flow of useful energy. If the maximum power principle is consistent with the optimization principle, then the two fields have developed convergent paradigms, which may lend credence to both. If they are mutually exclusive, then two possibilities exist. One of the paradigms must be incorrect, or there may be no emergent properties that define all systems. On the surface, the definitions of optimality and maximum power are identical. The maximization of currency intake by an organism is the same as the maximization of the flow of "useful" energy, assuming that any currency can be converted to energy equivalents. However, the two approaches appear to differ over two major issues. The resolution of these differences is a worthwhile goal. One major issue is control. Optimization principles treat control as unidirectional. Higher level subsystems constrain lower level systems. For ecosystems, Odum (1982) states that lower level subsystems (for example producers) will benefit 158 if they increase the feedback from higher-level subsystems (for example consumers). Here feedback is the flow of energy from higher- level to lower-level subsytems which influences the energy accrual properties of the lower-level subsystem. Odum (1982) has shown that this feedback mechanism will generally increase the ability of the lower-level subsystem to draw, or accrue, energy from a source. This viewpoint is reflected in the "maximum power" definition of net energy, which is the productive output (used oy higher subsystems) minus the feedback from higher subsystems. The "optimality" definition of net energy would be the intake of energy from lower systems minus the energy spent (or fed back) to acquire that intake. I define control as the capacity for constraint. In Chapter IV, decisions concerning time allocation set the boundary for, and therefore controlled, diet choice decisions. Here control is based only at the higher level. If control is generally based at all levels of a system, including control of higher-level subsystems by lower-level subsystems as implied above for ecosystems, then the modeling of behavioral decisions would be difficult without knowing the relationship between production and feedback. This relationship requires a model of the higher-level subsystem. Since there will always be a higher-level subsystem, multilevel control is an intractable problem in behavioral systems. On the other hand, if control is mediated through feedback from high-level systems to lower-level systems (as suggested by Mesarovic et al. 1970), then optimality and the maximum power principle may be functionally equivalent. Further work is clearly called for. A second issue that may differ between the approaches is the issue of embodied energy. Should the energetics of foraging be modeled using . JMTS"- 159 absolute or relative measures of currency? The answer lies in the use of the currency. If currency flow is compared between different subsystems, embodied energy may well be the most correct unit of comparison. However, if the flow or accrual rate of currency is to be determined through simulation or model dynamics, then absolute currency measures are required. Here embodied energy is implicit in the dynamics of the model. Hierarchical models of behavior, as presented here, involve a series of nested decision models about how the animal should behave. Each model is based on the appropriate currency or currencies. Of course, the setting up of a model does not mean that it will be of any use in the study of behavior. Before I discuss the role I think hierarchical models can play in the study of behavior, I will first review the contributions of earlier, non-hierarchical models. This is not intended to be a thorough literature review, since both optiraality and optimal foraging have been reviewed elsewhere (Cody 1974; Maynard Smith 1978; Pyke et al. 1977; Schoener 1971). I then discuss these examples in the light of the generalities we are attempting to draw from studies of foraging behavior. Non-Hierarchical Foraging Models There are numerous types of non-hierarchical models of foraging behavior, each of which addresses a specific aspect of foraging. Pyke et al. (1977) suggest that most optimal foraging models fall into four major categories: (I) choice of which food types to eat, (2) choice of 160 patch type to forage in, (3) allocation of time to different patches, and (4) patterns and rates of movement while foraging. In their review of optimal foraging studies, Pyke et al. (1977) note that these four decisions are assumed to be independent. However, the categories are not independent since they naturally fall under a hierarchical organization (Gass and Montgomerie 1981; Hassel and Southwood 1978; Charnov and Orians 1973)- ?or example, decisions about patch choice should reflect the type of prey expected to be taken from that patch (diet choice) and the amount of time that will be allocated to the patch. Thus all categories can be treated as a single, hierarchical model. All four categories listed above have been treated in the literature in a similar manner, so I will focus on one of the four categories, diet choice. The most widely used diet choice model is a variation of MacArthur and Pianka's (1966) original optimal foraging model. As Pyke et al. (1977) note, the same basic model has been (independently) derived by at least nine authors (Schoener 1969, 1971; MacArthur 1972; Charnov ■f976b; Charnov and Orians 1973; Timin 1973; Maynard Smith 1974; Pearson 1974; Pulliam 1974; Werner and Hall 1974; Estabrook and Dunham 1976). The diet choice model makes the following predictions (see Chapter IV for a more detailed discussion of Charnov' s [1976b] version): (I) prey should be ranked according to the ratio of net energy derived from the prey divided by the time required to handle that prey, (2) the forager should never exhibit partial prey preference (i.e., prey types that are a part of the diet should be taken whenever they are encountered), (3) the inclusion of any given prey type to the diet should be independent of the encounter rate of that prey type. 161 Prey types should be added to the diet until the net energy per unit handling time from a prey type falls below the net energy per unit foraging time derived from a diet of all higher ranking prey. The models were tested (mostly under laboratory conditions) on a number of different predators (Werner and Hall 1974; Krebs et al. 1977; Charnov 1976b; Zach and Falls 1978; Dunstone and O'Connor 1979; Hughes 1979; Palmer 1981; Goss-Custard 1977) and were found to be generally consistent with the predictions, although partial prey preference appeared to be exhibited more frequently than expected. ¥hat do these studies tell us? The studies cited above illustrate a number of features about foraging behavior. They suggest that the optiraality approach can be useful in studying diet choice under simple conditions. Thus foraging behavior can be treated as a selected trait, just as physiological (e.g., see HcNab 1980) or morphological (e.g., see Alexander 1982) traits can be. This is an important step because it allows us to use optimality to understand behavioral adaptations. The studies also give us some insight into the decision-making processes of animals, which is a primary focus of this field of research. The fact that the maximization of net energy appears to be a useful maximization parameter is also important because it suggests that energy may be a good parameter to start with in future studies. This by no means implies that it will work in all cases, but the utility of this maximization parameter can be tested. The fact that so many researchers have come up with the same model independently, and that the model appears to predict, at least qualitatively, the behavior of a variety of foragers, from insects to 162 mammals, suggests that the predictions are robust and have a high degree of generality. However, several studies have shown that the predictions will not hold for certain systems. In the following examples, each factor will change at least one of the predictions: (l) learning will change handling times and capture efficiencies (Murdoch 1969; McNair I98O), (2) foragers nay need to test habitats to monitor prey availability (Xrebs et al. 1977), (3) foragers often require some finite time to recognize prey (Charnov and Orians 1973; Hughes 1979), (4) a forager's diet choice may be constrained by nutrient requirements or toxicity of food (Pulliam 1975; Rapport 1981; Janzen and Freeland 1974), (5) the presence of predators may influence diet choice (wilinski and Heller 1978; Sih 1980b; Lucas 1983), and (6) foraging decisions nay be influenced by variability in prey quality or patch q^uality (Real 1980; Caraco 1979, 1980). These studies suggest that the list of factors originally included in the first diet choice models must be expanded. I should point out that the optimality approach was used to generate these new factors, and thus provided additional insights into the factors that influence foraging behavior. The studies listed above suggest that the original models are inappropriate under certain circumstances, but not that the foragers are foraging non-optiraally. Thus the original formulations will work under some circumstances and not under others. How do we treat these additional factors? They could be treated as special cases, and therefore we might conclude that the original models can produce robust predictions, but that the predictions are invalid in the special cases. On the other hand, we can add the new factors to an overall model for which there are no special cases. I suggest the latter approach. We should be able to generate a model that 163 embodies the essence of foraging behavior under a wide variety of conditions. This model should be far more predictive than any simple model of specific cases of foraging behavior. The best type of model that could be used is a hierarchical one. Hierarchical models will include the generality that is required and the flexibility needed to study foraging behavior in different sorts of animals and under different conditions. The original models were useful in their simplicity; this feature should not be discarded. Using hierarchical models, the simple models can be treated as special cases of the overall model. As a result, simple models can be used under the conditions for which these models are appropriate. Hierarchy and Optimality Models The lack of generality in the original diet choice models could have been predicted, given the simplicity of the models. This point brings up a slightly different issue: where should we look for generality in the first place? The approach to all optimality problems is to derive cost/benefit functions that are appropriate to the currency under consideration (Pyke et al. 1977). Thus, one generality of all optimal foraging models is that any given system can be studied using a cost/benefit function. Unfortunately, this generality does not go very far, since it makes no predictions about the type of decisions we should expect to see from a forager. The fact that the diet choice model was used under so many conditions and for so many types of organisms, suggests that researchers can use these models as general predictors of 164 behavior across a diversity of animal groups. The same can be said for the use of the marginal value theorem and its derivatives in the study of patch choice (Charnov i976a; Parker and Stuart 1976; Parker 1978; Sih 1980a; Cook and Gockrell 1978; Giller 1980; Orians and Pearson 1979; Krebs 1978; Cowie 1977). However, the work presented in Chapters II, III, and IV suggests that the diet choice models and the marginal value theorem may be much less general than was once thought. In some cases, the predictions may have been so general, especially when treated qualitatively, that they were upheld even when the systems were clearly inappropriate for the models. Partial prey consumption is a case in point (see Chapter II ) . This leads us back to the same question: What models will give predictions that are generally applicable to foraging behavior and that will therefore aid in a broad understanding of this behavior? The generality must enable us to gain insight about behavior, and not lead us to false conclusions. If the predictions generated from a model are so general that they cannot be rejected, even if the model is inappropriate, then the model has no power and should be discarded. The generalities we seek may be more in the framework or hierarchy than in the specific models. This can be illustrated using an excellent example from sex ratio theory. Sex ratio theory is an attempt to predict the variance in sex ratio patterns seen in nature. In fact, sex ratios have been shown to range from strongly male biased to strongly female biased. Sex ratios should reflect differences in investment pattern required to produce sons and daughters. Fisher (1958) originally proposed that the sex ratio produced by a female should be the inverse of the investment ratio. Sex ratios should be nroduced such i-i\w-jir.'' 165 that the female invests equally in both males and females. This theory- seems to predict the sex ratio pattern in a nur.ber of groups, from insects to mammals (Frank 1983; Charnov 1982). However, under certain demographic conditions, the sex ratio is more female biased than predicted. The fitness of a female within a deme is generally highest if that female invests equally in males and females (assuming no local mate competition; Hamilton 1967). However, the fitness between demes within a population is generally higher if the females in that deme produce mostly females, since demes with a female-biased sex ratio will produce more of the dispersing female offspring than a deme whose sex ratio is 50 percent males. Thus, there is a difference between the sex ratio patterns expected from the two levels of selection, within deme and between demes. If deme size is large enough, the effect of the lowest level will override the effect at the higher level and the sex ratio will be explained by Fisher's equal investment theory (again assuming no local mate competition). When deme size is very small and little breeding occurs between demes, the effect of inter-demic factors becomes more important and the sex ratio becomes more female biased. Conflicts between levels may also arise when subgenomic elements such as mitochondria or viral infections influence sex ratio. Here there may be a conflict between the equilibrium sex ratio at the level of the subgenomic element and the equilibrium sex ratio at the level of the individual. Which model produces the most generality? One could argue that each model (equal investment, group selection, and subgenomic selection) will work under specific conditions, and therefore will be generally applicable under those conditions. On the other hand, if each of the levels of selection is incorporated into a hierarchical 166 frame-rfork, then the framework itseli" will te generally applicable under all conditions. if there are no subgenomic particles and dene sise is large, the framework collapses to the original investment ratio theor;;-. If deme size is small and interdemic genetic exchange is low, the framework collapses to the group selection model. This feat^^re of ccllapsibility is especially compelling, since it may allow us to examine specific levels somewhat independently of other levels. Collapsibility vrill be most applicable when the model is collapsed on non-fesdback subsystems (see Hierarchy discussion above). ?his is because the constraints imposed by higher levels on the non-feedback subsystem can be treated as a constant in the model. Also the effect of decisions at lower levels are not incorporated into the decision process, therefore complex time components associated with feedback loops are lacking. Chapter IV is a good example of this. In this chapter I showed that diet choice should reflect the degree to which foraging bout length is constrained. The time constraints model (see Chapter IV) can be used irrespective of the factors that constrain bout length. Thus, the framework could be collapsed and treated as a single constraint within which diet choice decisions must be made. I cited studies where tidal cycle, predator pressure, and intra-specif ic pressures constrain bout length, ;/et the influence of each factor on diet choice is identical: as foraging bout length decreases, the diet should become more catholic. The regulation of foraging bout length is a decision about time sharing: how much time should the forager devote to each category of behavior (i.e. foraging, mating, etc.)? A study of this decision would be more complex, because time sharing is a feedback subsystem. To study this behavior properly, the dynamics of at least ■ -"A 167 two levels of decision need to te modeled. For example, ^-ith the time-sharing decisions wrens make when foraging for nestlings, the wren must decide to forage or stay and provide predator defense for her young. The two factors are not mutually exclusive, since feeding the young provides a predator deterrent. This level of decision was treated as a "black box" in my study. This decision is a feedback subsystem, because the amount of time required for foraging should be regulated by the net benefit derived from that foraging time. Thus, the study of time sharing in wrens requires at least two submodels to characterize the system. It may need more depending on the structure of the hierarchy. Most behavioral systems can be characterized by a hierarchy of decisions or effects. If we want to study the system as a whole, then we must study the entire hierarchy if we are to understand the behavior. As illustrated above, if we are interested in specific portions of the hierarchy, then the framework can be collapsed, depending on the type of subsystems characterizing the decision of interest. In chapter I, I showed that the structure of an antlicn's pit appears to maximize capture probability. The morphology of the pit included a layer of fine sand on the pit walls, which was shown to retard the escape of prey. Although there are a number of physical characteristics of sand that would tend to generate this type of design, the antlion clearly exhibits a number of behaviors that tend to make use of this particle-size/prey-escape relationship. Antlions regulate both particle velocity and trajectory angle. They also retain small particles inside the pit until the basic shape is attained, after which they line the pit with this fine sand. In addition to particle size distribution, the pit 168 is also characterized by two other features, slope and diameter. Particle size distribution is non-hierarchical in that there are no constraints imposed on this characteristic of pit morphology and the functioning of the pit. The more fine sand on the walls, the better it will function. Pit diameter is quite different. Here the energy invested in enlarging the pit should reflect the expected gain from that investment. The decision is characterized by a feedback subsystem. The decision cannot be addressed without a full knowledge of both the energetics of pit construction and the return from that investment. The same may hold true for pit slope. A simple study of pit morphology is possible due to the non-feedback nature of the decisions associated with that level of organization. A study of the regulation of pit diameter and slope will require a more thorough study to yield the same answer. The decision structure associated with pit-construction behavior would have to address the entire hierarchy, from pit morphology to pit diameter and its relationship with net benefit. When testing any model, the model gives us a framework within which we can view a system. The model also directs the experimental approach that is taken to study the system. In studies of foraging behavior, the specific models that have been formulated have been useful in some studies, but misleading in others. These mistakes appear to be caused be the restrictive nature of the single model approach. The use of systems-oriented, hierarchical models appears to be a much more powerful method of analysis. Once the system has been characterized, the hierarchy can be collapsed to address specific features of the system. Thus, a systems-level approach will enhance generality, since it can be used to address a broad spectrum of problems. It should increase our 169 ability to understand any system because it forces us to view a number of features of the system under study. It •/fill also allow us to simplify the question. This reduces the complexity of the question and increases our ability to measure the parameters that are important in each subsystem. 170 CONCLUSIONS In evaluating the utility of optimality modeling, two issues must be addressed. The first is whether or not the basic assumptions associated with the technique are consistent with observations. Of course, the most basic assumption is that animals are capable of making decisions that approach the theoretical optima. The second issue is a heuristic one. Does the technique provide insight into animal behavior that would otherwise be overlooked using other approaches? This dissertation covers three different temporal stages of foraging behavior, (l) the preparation for prey capture (specifically trap construction by antlion larvae), (2) the choice of prey, and (3) handling of prey once diet choice decisions have been made (specifically partial prey consumption). Optimal foraging models were used to evaluate each aspect. The results of the study indicate that foraging decisions are invariably constrained by factors that are unique to each stage of foraging. In each case when the constraints were built into foraging decision models, the behavior exhibited by the organisms under study was consistent with the predictions from the models. Trap-constructing behavior is consistent with the prediction that antlions should maximize the net energetic return from foraging through the manipulation of trap design. However, trap design is constrained by the physical properties of the material with which it is built. Thus the study of trap-constructing behavior can only be accomplished through 171 the characterization of physical constraints imposed on this behavior. Diet choice can be similarly constrained by features of the animal's environment. In systems where foraging time is limited through either biotic or abiotic factors, optimal diet choice must be modeled by explicitly including these factors in the model- Predictions from such models agree with the observed diet choice of a variety of animals. Time constraints may also be important in decisions concerning the utilization of prey by foragers that only partially consume prey. Predation modes (e.g. search or ambush methods) will impose unique constraints on the ability of animals to utilize their prey. Here again, if these constraints are included in foraging decision models, the output of the models agrees with the behavior exhibited by foraging animals. From the studies reported in this dissertation, it can be concluded that observations of animal behavior have been consistent with the assumptions and predictions from optimality models, but only if constraints are explicitly included in the models. Several models were reviewed in this dissertation which failed to incorporate the factors listed above. As a result, the authors of these studies concluded that the animals used in the experiments were not foraging optimally. If we ignore the fact that this conclusion is inappropriate to the technique (see Introduction and Chapter V), the inability of these earlier optimal foraging models to predict behavior suggests that these models were not robust enough to use under a diversity of foraging conditions. Since any given model provides a framework within which the researcher can formulate predictions, these models seem to restrict the perception of behavioral systems as much as they provide insight into behavioral adaptations. The use of optimality *5^«r" 172 models to study animal behavior appears to be valid, but the application of this technique to f oraging-behavior studies appears to be limiting our perception of behavioral systems. The non-hierarchical nature of most optimal foraging models is undoubtedly a major contributing factor in this limitation. I propose that models of behavior should be constructed using a hierarchical design. Unlike previous foraging models, hierarchical models require a synoptic perception of behavioral systems, and therefore are not restrictive in their predictive value. The utility of this approach is demonstrated in Chapter IV, where a single model was used to analyze foraging behavior of three taxonomically disparate organisms under extremely different environmental constraints. LITERATURE CITED Alexander, S.McN. 1982. Optima for animals. Arnold, London. Baker, M.C., C.S. Belcher, L.C. Deutsch, G.L. Sherman, and C.B. Thompson. 1981. Foraging success in junco flocks and the effects of social hierarchy. Anim. Behav. 29:137-142. Barnard, C.J. 1980. Equilibrium flock size and factors affecting arrival and departure in feeding house sparrows. Anim. Behav. 28:503-511 . Barr, A.J., Goodnight, J.H., Sail, J. P., Blair, W.H. and Chilko, D.M. 1979. SAS user's guide, 1979 edition. SAS Inst. Inc., Raleigh. Belovsky, G.E. 1981. Food plant selection by a generalist herbivore: the moose. Ecology 62:1020-1030. Bent, A.C. 1964. Life histories of north american nuthatches, wrens, thrashers, and their allies. Dover, New, York. Bernstein, B.B. and N. Jung. 1979. Selective pressures and coevolution in a kelp canopy community in southern California. Ecol. Monogr. 49:335-355. Bird, R.B., Stewart, W.E. and Lightfoot, E.N. I960. Transport phenomena. Wiley, New York. Brockmann, H.J., A. Grafen, and R. Dawkins. 1979- Evolutionarily stable nesting strategy in a digger wasp. J. Theor. Biol. 77:473-496. Caraco, T. 1979. Time budgeting and group size: a test of theory. Ecology 60:618-627. 1980. On foraging time allocation in a stochastic environment. Ecology 61 :119-128. Charnov, E. L. 1976a. Optimal foraging, the marginal value theorem. Theor. Popul. Biol. 9:129-136. 1Q76b. Optimal foraging: attack strategy of a mantid. Am. Nat. 110:141-151 . 173 15^- 174 1982. The theory of sex allocation. Princeton Univ., Princeton. and 3.H. Orians. 1973. Optimal foraging: some theoretical explorations. Charnov Press, Salt Lake. Cody, M.L. 1974. Optimization in ecology. Science 183:1156-1164. Cook, R. M. and B. J. Cockrell. 1978. Predator ingestion rate and its bearing on feeding time and the theory of optimal diets. J. Anim. Scol. 47:529-547- Cowie, R.J. 1977. Optimal foraging in great tits (Parus major). Nature 268:137-139- Cummins, K.W. 1967. Calorific equivalents for studies in ecological energetics. Pymatuning Laboratory in Ecology, Univ. Pittsburgh, Pittsburgh. Dawkins, R. 1976. Hierarchical organisation: a candidate principle for ethology, p 7-54 in P.P.G. Bateson and R.A. Hinde, eds. Growing points in ethology. Cambridge Press, Cambridge. DeBenedictis, P. A., F. B. Gill, F. R. Hainesworth, G. Pyke and L.L. yolfe. 1978. Optimal meal size in hummingbirds. Am. Hat. 112:301-316. DeJours, P. 1975- Principles of comparative respiratory physiology- North-Holland, Amsterdam- Denny, M. 1976. The physical properties of spider's silk and their role in the design of orb-webs. J. Exp. Biol. 65:483-506. Dunstone, N. , and R.J. 0' Conner. 1979- Optimal foraging in an amphibious mammal. I. The aqualung effect. Anim. Behav. 27:1182-1194. Eberhard, W.G. 1980. Spider and fly play cat and mouse. Nat. Hist. 89:56-61 . Elner, R.W. , and R.N. Hughes. 1978. Energy maximization in the diet of the shore crab, Carcinus maenas. J. Anim. Ecol. 47:103-116. Emlen, J.M. 1966. The role of time and energy in food preference. Am. Nat. 100:611-617. Estabrook, G.F., and A.E. Dunham. 1976. Optimal diet as a function of absolute abundance, relative abundance, and relative value of available prey. Am. Nat. 110:401-413- Fisher, R.A. 1958. The genetical theory of natural selection. Dover, New York. 175 } -l Ford, M.J. 1977. Energy costs of the predation strategy of the -^eb-spinning spider Lepthyphantes zimmerTnanni 3ertkau (Linyphiidae). Oecologia 28: 341 -349- Frank, S. 1983. A hierarchical view of sex ratio patterns. Fla. Entomol. (in press). Freed, L.A. 1981. Optimal foraging by house wrens in the face of conflicting demands. Anim. Behav. Soc. Ann. Meeting Abst. Gass, C.L., and R.D. Montgomerie. 1981. Hummingbird foraging behavior: decision making and energy regulation, p 159-194 in A.C. Kamil and T.D. Sargent, eds. Foraging behavior: ecological, ethological, and psychological approaches. Garland, New York. Giller, P.S. 1980. The control of handling time and its effects on the foraging strategy of a heteropteran predator, Notonecta. J. Anim. Ecol. 49:699-712. Gittelman, S. H. 1974. Locomotion and predatory strategy in '-' backswimmers (Hemiptera: Notonectidae) . Am. Midi. Nat. 92:496-500. Goss-Custard, J.D. 1977. Optimal foraging and the size selection of '■ . worms by redshank (Tringa totanus) in the field. Anim. Behav. , , 25:10-29. ■:'■ ;• 1981. Feeding behavior of redshank, Tringa totanus, and optimal foraging theory, p 115-134 in A.C. Xamil and T.D. Sargent, eds. Foraging behavior: ecological, ethological, and psychological approaches. Garland, New York. Gould, S.J. 1982. The meaning of punctuated equilibrium and its role in validating a hierarchical approach to macroevolution. p 83-104, in R. Milkman, ed. Perspectives on evolution. Sinauer, Sutherland. Greenstone, M.H. 1979- Spider feeding behavior optimises dietary essential amino acid composition. Nature 282:501-503- Griffiths, D. 1980. The feeding biology of ant-lion larvae: prey capture, handling, and utilization. J. Anim. Ecol. 49:99-125- 1982. Tests of alternative models of prey consumption by predators, using ant-lion larvae. J. Anim. Ecol. 51:363-374. Hamilton, V.D. 1967. Extraordinary sex ratios. Science 156:477-488. Hassell, M. P., J. H. Lawton and J- R. Beddington. 1976. The components of arthropod predation. I. The prey death rate. J. Anim. Ecol. 45:135-164. 176 Hassel, M.P., and T.H.S. Southwood. 1976. Foraging strategies of insects. Ann. Rev. Ecol. Syst. 9:75-98. Hemmingsen, A.M. I960. Energy metabolism as related to body size and respiratory surfaces and its evolution. Rep. Steno Mem. Hospital, Nordisk Insulinlaboratorium. Vol IX. Hervey, G.R. 1969. Regulation of energy balance. Nature 222:529-631. Hieber, C.S. 1979. Web orientations to wind and light in the laboratory for the spiders Araneus diadematus Clerck and Araneus gemmoides Chamberlin and Ivie (Araneae, Araeidae). M.S. thesis, University of North Dakota. Hildrew, A.G. and Townsend, C.R. 1980. Aggregation, interference and foraging by larvae of Plectrocnemia conspersa (Trichoptera: Polycentropodidae) . Anim. Behav. 28:553-560. Houston, A.I., T. Halliday and D. McFarland. 1977. Towards a model of the courtship of the smooth newt, Triturus vulgaris, with special reference to problems of observability in the stimulation of behaviour. Med. Biol. Engng. 15:49-61. Moiling, C. S. 1965. The functional response of predators to prey density and its role in mimicry and population regulation. Mem. Entomol. Soc. Canada 45:1- 60. 1956. The functional response of invertebrate predators to prey density. Mem. Entomol. Soc. Canada 48:1-86. Hughes, R. N. 1979- Optimal diets under the energy maximization premise: the effects of recognition time and learning. Am. Nat. 113:209-21 1 . Jaeger, R.G., and D.E. Barnard. 1981. Foraging tactics of a terrestrial salamander: choice of diet in structurally simple environments. Am. Nat. 117:639-664. , H.G. Joseph, and D.E. Barnard. 1981. Foraging tactics of a terrestrial salamander: sustained yield in territories. Anim. Behav. 29:1100-1105. Janzen, D.H. and W.J. Freeland. 1974. Strategies in herbivory by mammals: the role of plant secondary compounds. Am. Nat. 108:269-289. Johnson, D. M. , B. G. Akre and P. H. Crowley. 1975. Modeling arthropod predation: wasteful killing by damselfly naiads. Ecology 56:1081-1093. Kaplan, W. and D.J. Lewis. 1971. Calculus and linear algebra. Wiley, New York. -«' 177 Krebs , J.R. 1978. Optimal foraging: decision rules for predators, p 23-62 _iri J.R. Krebs and N.B. Davies, eds. Behavioural ecology: an evolutionary approach. Blackwell, Oxford. ^ j.T. Erichsen, M.J. Webber, and E.L. Charnov. 1977. Optimal prey selection in the great tit (Parus major). Anim. Behav. 25:30-38. Krynine, D. 1941. Soil mechanics. McGraw-Hill, New York. Lewontin, R.C. 1970. The units of selection. Ann. Rev. Ecol. Syst. 1 :1-13. Lucas, J.R. 1982. The biophysics of pit construction by antlion larvae (Myrmeleon, Neuroptera). Anim. Behav. 30:651-664. 1983. The role of foraging time constraints and variable prey encounter in optimal diet choice. Am. Nat. (in press). and Brockmann, H.J. 1981. Predatory interactions between ants and antlions (Hymenoptera: Formicidae and Neuroptera: Myrmeleontidae) . J. Kansas Entom. Soc. 54:228-232. and L.A. Stange. 1981. Key and descriptions to the Myrmeleon larvae of Florida (Neuroptera: Myrmeleontidae). Fla. Entomol. 64:207-216. MacArthur, R.H. 1972. Geographical ecology. Harper and Row, New York. and E.R. Pianka. 1966. On optimal use of a patchy environment. Am. Nat. 100:603-609. Marachi, N.D., Chan, O.K. and Seed, H.B. 1972. Evaluation of properties of rockfill materials. J. Soil Mech. Found. Div., Proc. Am. Soc. Civil Engng. 98:95-114. Maynard Smith, J. 1974. Models in ecology. Cambridge Press, Cambridge. 1978. Optimization theory in evolution. Ann. Rev. Ecol. Syst. 9:31-56. Mayzaud, P. and S. A. Poulet. 1978. The importance of the time factor in the response of zooplankton to varying concentrations of naturally occurring particulate matter. Limnol. Oceanogr. 23:1144-1154. McCleery, R. 1978. Optimal behaviour sequences and decision making, p 377-410, in J.R. Krebs and N.B. Davies, eds. Behavioural ecology: an evolutionary approach. Blackwell, Oxford. McClure, M.S. 1976. Spatial distribution of pit-making antlion larvae (Neuroptera: Myrmeleontidae): density effects. Biotropica 8:179-183. 178 McFarland, D.J. 1977. Decision making in ani.-nals. Nature 269:15-21. McNab, B.K. 19S0. Food habits, energetics, and the population biology of mammals. Am. Nat. 116:106-124. McNair, J.N. 1980. A stochastic foraging model with predator training effects: I. Functional response, switching, and run lengths. Theoret. Pop. Biol. 17:141-166. Menge, J.L. 1974. Prey selection and foraging period of the predaceous rocky intertidal snail, Acanthina punctulata. Oecologia 17:293-316. Mesarovic, M.D., D. Macko , and Y. Takahara. 1970. Theory of hierarchical, multilevel, systems. Academic Press, New York. Milinski, M. , and R. Heller. 1978. Influence of a predator on the optimal foraging behavior of sticklebacks (Gasterosteus aculeatus) . Nature 275:642-644. Murdoch, W.W. 1969. Switching in general predators: experiments on predator specificity and stability of prey populations. Ecol. Monogr. 39:335-354. , S. Avery, and M.E.B. Smyth. 1975. Switching in predatory fish. Ecology 56:1094-1105. Nakamura, K. 1977. A model for the functional response of a predator to varying prey densities; based on the feeding ecology of wolf spiders. Bull. Natl. Inst. Agric. Sci. 31:29-89. Norberg, R.A. 1981. Temporary weight decrease in breeding birds may result in more fledged young. Am. Nat. 118:838-350. Oaten, A., and W.W. Murdoch. 1975. Switching, functional response, and stability in predator-prey systems. Am. Nat. 109:299-318. Odum, H.T. 1983. Systems ecology: an introduction. Wiley, New York. Olmstead, L.R., s. Krater, G.E. Williams, and R.G. Jaeger. 1979. Foraging tactics of the mimic shiner in a two-prey system. Copeia 1979:437-441 . Orians, G.K. and N.P. Pearson. 1979. On the theory of central place foraging. p 1 55-1 77 in D.H. Horn, R. Mitchell, and G.R. Stairs, eds. Analysis of ecological systems. Ohio St. Univ., Columbus. Owen-Smith, N. and P. Novellie. 1982. What should a clever ungulate eat? Am. Nat. 119:151-178. 179 Palmer, A.R. 1981. Predatory errors, foraging in unpredictable environments and risk: the consequences of prey variation in handling time versus net energy. Am. Nat. 118:908-915. Parker, G.A. 1978. Searching for mates, p 214-244 in J.R. Xrebs and N.B. Davies, eds. Behavioural ecology: an evolutionary approach. Blackwell, Oxford. and R. A. Stuart. 1975. Animal behavior as a strategy optimizer: evolution of resource assessment strategies and optimal emigration thresholds. Am. Nat. 110:1055-1075. Pearson, N.E. 1974. Optimal foraging theory. Quan. Science Paper 39- Center for Quant. Science in Forestry, Fisheries and Wildlife, Univ. Washington, Seattle. Powell, G.V.N. 1974. Experimental analysis of the social value of flocking by starlings (Sturnus vulgaris) in relation to predation and foraging. Anim. Behav. 22:501-505. Prestwich, K.N. 1977- The energetics of web-building in spiders. Comp. Biochem. Physiol. 57A:321-526. Pulliam, H.R. 1974. On the theory of optimal diets. Am. Nat. 108:59-74. 1975. Diet optimization with nutrient constraints. Am. Nat. 109:765-768. Pyke, G. H. 1978. Optimal foraging: movement patterns of bumblebees between inflorescences. Theor. Pop. Biol. 13:72-98. , H.R. Pulliam, and E.L. Charnov. 1977. Optimal foraging: a selective review of theory and tests. Quart. Rev. Biol. 52:137-154. Rapport, D.J. 1971. An optimization model of food selection. Am. Nat. 105:575-587. 1980. Optimal foraging for complementary resources. Am. Nat. 116:324-346. 1981. Foraging behavior of Stentor coerulus: a microeconomic interpretation, p 77-94 in A.C. Karail and T.D. Sargent, eds. Foraging behavior: ecological, ethological, and psychological approaches. Garland, New York. Real, L.A. 1980. On uncertainty and the law of diminishing returns in evolution and behavior. p 37-64 in J.E.R. Staddon, ed. Limits to action: the allocation of individual behavior. Academic, New York. 180 Sandness J. N. and J. A. McMurtry. 1970. Functional response of three species of phytoseiidae (Acarira) to prey density. Can. Entomol. 102:692-704. Schluter, D. 1981. Does the theory of optimal diets apply in complex environnients? Am. Nat. 113:139-147. Schoener, T.W. 1963- Models of optimal size for a solitary predator. Am. Nat. 103:277-313. 1971. Theory of feeding strategies. Ann. Rev. Ecol. Syst. 2:369-404. Shoemaker, C. A. 1977. Mathematical construction of ecological models, p 75-114 _in C.A.S. Hall and J.W. Day, eds. Ecological modeling in theory and practice: an introduction with case histories. Wiley, New York. Sibly, R. and D.J. McFarland. 1975. On the fitness of behavior sequences. Am. Nat. 110:601-617. Siegel, S. 1956. Nonparametric statistics for the behavioral sciences. McGraw-Hill, New York. Sih, A. 1 980a_. Optimal foraging: partial consumption of prey. Am. Nat. 116:281-290. 1980b. Optimal behavior: can foragers balance two conflicting demands? Science 21 0: 1041 -1043- Simon, H. A. 1956. Rational choice and the structure of the environment. Psychol. Rev. 2:129-138. Singh, A. 1976. Soil engineering in theory and practice. Vol 1, 2nd edn. Asia Publishing House, New York. Sokal, R.R. and F.J. Rohlf. 1969. Biometry. Freeman, San Francisco. Timin, M.S. 1973- A multi-species consumption model- Math. Biosciences 16:59-66. Tinbergen, N. 1969. The study of instinct. Oxford, New York. Topoff, H. 1977. The pit and the antlion. Nat. Hist. 86:65-71. Treherne, J. E. and V. A. Foster. 1981. Group transmission of predator avoidance behaviour in a marine insect: the Trafalgar effect. Anim. Behav. 29:911-917. Turnbull, A. 1973. Ecology of the true spiders (Araeneomorphae) . Ann. Rev. Ent. 18:305-348. i^isfh'-^ 181 Turner, C.H. 1915. Notes on the behavior of the ant-lion with emphasis on the feeding activities anr" letisimulation. Biol. Bull. 29:277-307. Tusculescu, R., Topoff, H. and Wolfe, S. 1975. Mechanisms of pit construction by antlion larvae. Ann. Ent. Soc. Amer. 68:719-720. Uetz, G.W., Johnson, A.D. and Schemske.D.W. 1978. Web placement, web structure and prey capture in orb-weaving spiders. Bull. Br. Arachnol. Soc. 4:141-148. van der Kloot, W.G. 1960. Neurosecretion in insects. Ann. Rev. Entomol. 5:35-52. von Frisch, K. 1974. Animal architecture. Harcourt Brace Jovanovich, New York. Werner, E.E. and D.J. Hall. 1974. Optimal foraging and size selection of prey by the bluegill sunfish (Lepomis machrochirus) . Ecology 55:1042-1052. Vestoby, M. 1974. An analysis of diet selection by large generalist herbivores. Am. Nat. 108:290-304. 1978. What are the biological bases of varied diets? Am. Nat. 112:627-631 . Wheeler, W.M. 1930. Demons of the dust. W.W. Norton, New York. Wilson, D.S. 1974. Prey capture and competition in the antlion. Biotropica 6:187-193. Witt, P.N., Rawlings, J.O. and Reed, C.F. 1972. Ontogeny of web-building behavior in two orb-weaving spiders. Am. Zool. 12:445-454. , Reed, C.F. and Peakall, D.B. 1968. A spider's web: problems in regulatory biology. Springer-Verlag, New York. Work, R.W. 1981. Web components associated with the major ampullate silk fibers of orb-web-building spiders. Trans. Am. Microsc. Soc. 100:1-20, Youthed, G.J. and Moran, V.C. 1969. Pit construction by myrmeleontid larvae. J. Insect Physiol. 15:867-875- Zach, R. and J.B. Falls. 1978. Prey selection by captive ovenbirds (Aves: Parulidae). J. Anim. Ecol. 47:929-943. and J.N.M. Smith. 1981. Optimal foraging in wild birds? p 95-109 in A.C. Kamil and T.D. Sargent, eds. Foraging behavior: ecological, ethological, and psychological approaches. Garland, New York. BIOGRAPHICAL SKETCH Jeffrey Robert Lucas was born in Januarj-, 1953. He received his B.S. degree at the Florida Institute of Technology and his M.S. degree at the University of Florida. During his tenure at the University of Florida, Jeff worked on a project funded through the Florida Power Corporation which evaluated the effect of heated effluents on the productivity of estuarine bays. He also worked as a teaching assistant in the Department of Zoology. The fact that he met and married Lynda Peterson should not be overlooked. He has received a NATO Postdoctoral Fellowship to continue his antlion project with Dr. John Krebs at Oxford. Tally ho. 182 I certify that I have read this study and that in my opinion it conforms to acceptable standards of scholarly presentation and is fully adequate, in scope and quality, as a dissertation for the degree of Doctor of Philosophy. H. Jane Brockmann, Chairperson Associate Professor of Zoology I certify that I have read this study and that in my opinion it conforms to acceptable standards of scholarly presentation and is fully adequate, in scope and quality, as a dissertation for the degree of Doctor of Philosophy. C^^t'V.-V^vvvt^ ll< f^^ Carmine Lanciani Professor of Zoology I certify that I have read this study and that in my opinion it conforms to acceptable standards of scholarly presentation and is fully adequate, in scope and quality, as a dissertation for the degree of Doctor of Philosophy. yjdiiUil ^ / Brian K. McNab Professor of Zoology I certify that I have read this study and that in my opinion it conforms to acceptable standards of scholarly presentation and is fully adequate, in scope and quality, as a dissertation for the degree of Doctor of Philosophy. Frank Nordtie Professor of Zoology I certify that I have read this study and that in my opini-n it conforms t: acceptable standards if scholarly presentation and is fully adequate, in scope and quality, as a dissertation for the degree of Doctor of Philosophy. -^CC"^^ Howard T. Odum Graduate Research Professor of Environmental Engineering Sciences This dissertation was submitted to the Graduate Faculty of the Department of Zoology in the College of Liberal Arts and Sciences and to the Graduate Council, and was accepted as partial fulfillment of the requirements for the degree of Doctor of Philosophy. April, 1983 Dean for Graduate Studies and Research & • n UNIVERSITY OF FLORIDA 3 1262 08554 0077 }