Personal tools
Home » Working Groups » EBM Marine Predators » Endo-Ecto Group » Predator functional responses: Discriminating between handling and digesting prey. Ecological Monographs 72 (1): 95-112
Document Actions

Predator functional responses: Discriminating between handling and digesting prey. Ecological Monographs 72 (1): 95-112

Keywords: consumer-resource systems; consumption rate; digestion-limited predators; digestion time; functional response models; handling-limited predators; handling time; hunger level; predation rate; predator–prey systems; steady-state satiation (SSS) equation. Abstract: We present a handy mechanistic functional response model that realistically incorporates handling (i.e., attacking and eating) and digesting prey. We briefly review current functional response theory and thereby demonstrate that such a model has been lacking so far. In our model, we treat digestion as a background process that does not prevent further foraging activities (i.e., searching and handling). Instead, we let the hunger level determine the probability that the predator searches for new prey. Additionally, our model takes into account time wasted through unsuccessful attacks. Since a main assumption of our model is that the predator’s hunger is in a steady state, we term it the steady-state satiation (SSS) equation. The SSS equation yields a new formula for the asymptotic maximum predation rate (i.e., asymptotic maximum number of prey eaten per unit time, for prey density approaching infinity). According to this formula, maximum predation rate is determined not by the sum of the time spent for handling and digesting prey, but solely by the larger of these two terms. As a consequence, predators can be categorized into two types: handling-limited predators (where maximum predation rate is limited by handling time) and digestion-limited predators (where maximum predation rate is limited by digestion time). We give examples of both predator types. Based on available data, we suggest that most predators are digestion limited. The SSS equation is a conceptual mechanistic model. Two possible applications of this model are that (1) it can be used to calculate the effects of changing predator or prey characteristics (e.g., defenses) on predation rate and (2) optimal foraging models based on the SSS equation are testable alternatives to other approaches. This may improve optimal foraging theory, since one of its major problems has been the lack of alternative models.
Ecological Monographs, 72(1), 2002, pp. 95–112
 2002 by the Ecological Society of America




   PREDATOR FUNCTIONAL RESPONSES: DISCRIMINATING BETWEEN
         HANDLING AND DIGESTING PREY
              JONATHAN M. JESCHKE,1,3 MICHAEL KOPP,2          RALPH TOLLRIAN1
                                        AND

     1Department of Ecology, Zoological Institute, Ludwig-Maximilians-Universitat Munchen, Karlstrasse 25,
                                           ¨  ¨
                        D-80333 Munchen, Germany
                              ¨
                             ¨
      2Max-Planck-Institut fur Limnologie, Abteilung Okophysiologie, Postfach 165, D-24302 Plon, Germany
                 ¨                                ¨

         Abstract. We present a handy mechanistic functional response model that realistically
       incorporates handling (i.e., attacking and eating) and digesting prey. We briefly review
       current functional response theory and thereby demonstrate that such a model has been
       lacking so far. In our model, we treat digestion as a background process that does not
       prevent further foraging activities (i.e., searching and handling). Instead, we let the hunger
       level determine the probability that the predator searches for new prey. Additionally, our
       model takes into account time wasted through unsuccessful attacks. Since a main assumption
       of our model is that the predator’s hunger is in a steady state, we term it the steady-state
       satiation (SSS) equation.
         The SSS equation yields a new formula for the asymptotic maximum predation rate
       (i.e., asymptotic maximum number of prey eaten per unit time, for prey density approaching
       infinity). According to this formula, maximum predation rate is determined not by the sum
       of the time spent for handling and digesting prey, but solely by the larger of these two
       terms. As a consequence, predators can be categorized into two types: handling-limited
       predators (where maximum predation rate is limited by handling time) and digestion-limited
       predators (where maximum predation rate is limited by digestion time). We give examples
       of both predator types. Based on available data, we suggest that most predators are digestion
       limited.
         The SSS equation is a conceptual mechanistic model. Two possible applications of this
       model are that (1) it can be used to calculate the effects of changing predator or prey
       characteristics (e.g., defenses) on predation rate and (2) optimal foraging models based on
       the SSS equation are testable alternatives to other approaches. This may improve optimal
       foraging theory, since one of its major problems has been the lack of alternative models.
        Key words: consumer-resource systems; consumption rate; digestion-limited predators; digestion
       time; functional response models; handling-limited predators; handling time; hunger level; predation
       rate; predator–prey systems; steady-state satiation (SSS) equation.


            INTRODUCTION                 of changing predator or prey characteristics (e.g., de-
                                   fenses) on predation rate.
  The relationship between predation rate (i.e., number
of prey eaten per predator per unit time) and prey den-            PREVIOUS MODELS: A BRIEF REVIEW
sity is termed the ‘‘functional response’’ (Solomon
                                     Scientists have been modeling functional responses
1949). It is specific for each predator–prey system. The
                                   since the 1920s (reviewed by Holling 1966, Royama
term predator is meant in its broadest sense here, i.e.,
                                   1971), although the term ‘‘functional response’’ was
it includes carnivores, herbivores, parasites, and par-
                                   only introduced in 1949 by Solomon. Since, to our
asitoids. The functional response is an important char-
                                   knowledge, the last review of functional response mod-
acteristic of predator–prey systems and an essential
                                   els dates back to 1971 (Royama), we provide an over-
component of predator–prey models: Multiplying the
                                   view of models published since 1959 together with the
functional response with predator population density
                                   most important factors incorporated in each model (Ta-
and a time factor yields the total number of prey eaten
                                   ble 1). In addition, Fig. 1 shows a ‘‘family tree’’ of
in the period of interest, e.g., one year or one prey
                                   these functional response models. Holling (1959a) has
generation. Given further information, such as actual
                                   categorized functional responses into three main types,
predator density and an energy conversion factor, one
                                   which he called type I, II, and III. Our discussion will
can assess future population densities of both predator
                                   focus on type II functional responses, since these have
and prey. With a mechanistic functional response mod-
                                   been most frequently observed (Hassell et al. 1976,
el, as presented in this study, one can predict the effects
                                   Begon et al. 1996). They are characterized by a hy-
                                   perbolic curve. Starting at low prey densities on the
  Manuscript received 8 May 2000; revised 14 December 2000;
                                   abscissa, predation rate first increases almost linearly
accepted 7 February 2001; final version received 12 March 2001.
                                   until it gradually slows down to reach an upper limit.
  3 E-mail: jonathan.jeschke@gmx.net


                                 95
                                                      Ecological Monographs
                       JONATHAN M. JESCHKE ET AL.
96
                                                          Vol. 72, No. 1

TABLE 1.  A selection of functional response models.

                                        Model
                     A    B   C    D   E    F   G    H    I   J    K    L
       Features         (C)   (C)  (CP)  (C)  (P)   (C)  (F)   (F)   (F)  (P)   (C)   (C)
Components
 Success rate1
 Probability of attack
 Handling time2
 Searching and handling overlap-
  ping3
 Hunger and satiation4
 Handling prey   digesting prey
 Adaptive behavior5
 Incomplete consumption6
 Nonforaging activities7
 Spatial heterogeneity8
 Temporal heterogeneity9
 Stochasticity10
 Environmental conditions11
 Predator injury by prey
 Inducible defenses12
 Dependent on prey density
  Prey density
  Decreasing prey density13
  Alternative prey14
  Learning or switching15
  Swarming effect16
 Dependent on predator density
  Predator density
  Interference between predators17
  Multiple predator effects18
Functional response types
  Type I
  Type II
  Type III
  Dome shaped
  Other forms19
  Notes: Small capital letters in parentheses under models indicate the kind of predator that the model was primarily designed
for: C, carnivores; F, filter feeders; H, herbivores, P, parasites or parasitoids. In the body of the table, ‘‘ ’’ means the model
includes that component, ‘‘ ’’ means the model additionally includes subcomponents, and ‘‘ ’’ means the model does not
include that component. Sources for models are as follows: (A) Gause (1934), Ivlev (1961), Eq. 1; (B) Rashevsky (1959;
no overall model but different equations); (C) Watt (1959); (D) Royama (1971: Eq. 3.12), see also Nakamura (1974: Eq.
15); (E) Royama (1971: Eq. 3.24); (F) Nakamura (1974); (G) Sjoberg (1980); (H) Lam and Frost (1976); (I) Lehman (1976);
                                  ¨
(J) Casas et al. (1993); (K) Disc equation (Holling 1959b), Eq. 2; (L) Invertebrate model (Holling 1966; see also Metz and
van Batenburg 1985a,b); (M) Vertebrate model (Holling 1965); (N) Holling and Buckingham (1976); (O) Rao and Kshiragar
(1978); (P) Metz et al. (1988; see also Metz and van Batenburg 1985a, b); (Q) Cushing (1968); (R) Tostowaryk (1972); (S)
Random predator equation (Royama 1971, Rogers 1972); (T) Random parasite equation (Royama 1971, Rogers 1972); (U)
Beddington (1975); (V) Hassell et al. (1977); (W) Longstaff (1980); (X) Mills (1982); (Y) Crowley (1973); (Z) Oaten and
Murdoch (1975); (AA) Real (1977); (BB) McNair (1980); (CC) Abrams (1982); (DD) Dunbrack and Giguere (1987); (EE)
Abrams (1990a); (FF) Descriptive equation (Fujii et al. 1986); (GG) Ungar and Noy-Meir (1988); (HH) Random patch model
         ˚¨
(Lundberg and Astrom 1990; see also Lundberg and Danell 1990); (II) Juliano (1989); (JJ) Fryxell (1991; see also Wilmshurst
et al. 1995, 1999, 2000); (KK) Spalinger and Hobbs (1992; see also Laca et al. 1994, Shipley et al. 1994); (LL) Farnsworth
and Illius (1996; see also Laca et al. 1994, Shipley et al. 1994); (MM) Hirakawa (1997b; see also Laca et al. 1994, Shipley
et al. 1994); (NN) Farnsworth and Illius (1998; see also Laca et al. 1994, Shipley et al. 1994); (OO) Ruxton and Gurney
(1994); (PP) Cosner et al. (1999) [This model closes a gap between density dependent and ratio dependent functional response
models. Purely ratio dependent models are not included in Table 1, but see Arditi and Ginzburg (1989). However, as Berryman
et al. (1995) have written: ‘‘Note that prey-dependent functional responses can be transformed into ratio-dependent functional
responses by substituting the prey/predator ratio for prey density in the equation.’’]; (QQ) Streams (1994); (RR) Schmitz
(1995; see also Abrams [1990c] and review by Schmitz et al. [1997]); (SS) Abrams and Schmitz (1999); (TT) Berec (2000;
see also Engen and Stenseth 1984); (UU) SSS equation (Eq. 13).
  1 Success rate consists of four subcomponents: (1) encounter rate, (2) probability of detection, (3) hunger-independent

probability of attack, and (4) efficiency of attack; empirical values for the attack efficiencies of predators have been reviewed
by Curio (1976), Vermeij (1982), and Packer and Ruttan (1988).
  2 Handling time (per prey item) includes attacking time (including evaluating, pursuing, and catching time) and eating

time. See also Anholt et al. (1987), Demment and Greenwood (1988), Laca et al. (1994), Parsons et al. (1994), and Shipley
et al. (1994).
  3 Important for queueing predators (Juliano 1989; see also Visser and Reinders 1981, Lucas 1985, Lucas and Grafen 1985)

and vertebrate herbivores (Spalinger and Hobbs 1992, Parsons et al. 1994, Farnsworth and Illius 1996, 1998, Hirakawa
1997b; see also Laca et al. 1994, Shipley et al. 1994).
                       PREDATOR FUNCTIONAL RESPONSES
Febraury 2002                                                          97

TABLE 1.    Extended.

                                 Model
 M    N    O    P   Q   R   S   T   U    V   W   X   Y   Z  AA   BB   CC  DD   EE
 (C)  (C)   (C)   (C)  (F)  (C)  (C)  (P)  (CP)  (CP)  (C)  (C)  (F)  (C)  (C)  (C)  (C)  (CF)  (C)




  4 Some models include predator satiation via a maximum predation rate determined by the characteristics of the digestive

system (‘‘ ’’). Other models include the fact that the predator’s gut content is increased by ingestion and decreased by
digestion (‘‘ ’’). See also Campling et al. (1961), Curio (1976), Belovsky (1978, 1984a, b, c, 1986a, b, 1987), Mayzaud and
Poulet (1978), Bernays and Simpson (1982), Murtaugh (1984), Crisp et al. (1985), Demment and Greenwood (1988), Verlinden
and Wiley (1989), Illius and Gordon (1991), Doucet and Fryxell (1993), Forchhammer and Boomsma (1995), Henson and
Hallam (1995), Hirakawa (1997a), and Wilmshurst et al. (2000).
  5 See also Belovsky (1978, 1984a, b, c, 1986a, b, 1987), Cook and Cockrell (1978), Sih (1980, 1984), Owen-Smith and

Novellie (1982), McNair (1983), Abrams (1984, 1987, 1989, 1990b, c, 1991, 1992, 1993), Engen and Stenseth (1984),
Formanowicz (1984), Lucas (1985), Wanink and Zwarts (1985), Stephens and Krebs (1986), Anholt et al. (1987), Demment
                                         ˚¨
and Greenwood (1988), Belovsky et al. (1989), Verlinden and Wiley (1989), Astrom et al. (1990), Lundberg and Danell
(1990), Mitchell and Brown (1990), Abrams and Matsuda (1993), Doucet and Fryxell (1993), Werner and Anholt (1993),
McNamara and Houston (1994), Forchhammer and Boomsma (1995), Hirakawa (1995, 1997a), Fryxell and Lundberg (1997),
Leonardsson and Johansson (1997), Rothley et al. (1997), Schmitz et al. (1997), and Wilmshurst et al. (2000).
  6 See also Buckner (1964), Johnson et al. (1975), Curio (1976), Cook and Cockrell (1978), Sih (1980), Owen-Smith and

Novellie (1982), McNair (1983), Formanowicz (1984), Lucas (1985), Lucas and Grafen (1985), Metz and van Batenburg
       ˚¨
(1985a, b), Astrom et al. (1990), Lundberg and Danell (1990), and Fryxell and Lundberg (1997).
  7 For example, avoidance of top predators, migration, molting, reproductive activities, resting, sleeping, territorial behavior,

thermoregulation, and times of slow rates of metabolism like winter dormancy; see also Belovsky (1978, 1984a, b, c, 1986a,
b, 1987), Caraco (1979), Herbers (1981), Bernays and Simpson (1982), Owen-Smith and Novellie (1982), Abrams (1984,
1991, 1993), Stephens and Krebs (1986), Belovsky et al. (1989), Verlinden and Wiley (1989), Bunnell and Harestad (1990),
Mitchell and Brown (1990), McNamara and Houston (1994), Forchhammer and Boomsma (1995), Hirakawa (1997a), Leon-
ardsson and Johansson (1997), and Rothley et al. (1997).
  8 See also Griffiths and Holling (1969), Paloheimo (1971a, b), Oaten (1977), May (1978), Real (1979), McNair (1983),

Belovsky et al. (1989), Blaine and DeAngelis (1997), Fryxell and Lundberg (1997), and Wilmshurst et al. (2000), among others.
  9 For example, diel or annual periodicity (Curio 1976, Bernays and Simpson 1982, Belovsky et al. 1989, Forchhammer

and Boomsma 1995).
  10 See also Paloheimo (1971a, b), Curry and DeMichele (1977), Curry and Feldman (1979), McNair (1983), Lucas (1985),

and Metz and van Batenburg (1985a, b).
  11 For example, precipitation, temperature (Fedorenko 1975, Thompson 1978, Bernays and Simpson 1982), and wind.
  12 Behavioral and morphological defenses, that are not permanently present but are induced by the predator (e.g., Fryxell

and Lundberg 1997, Karban and Baldwin 1997, Tollrian and Harvell 1999, Jeschke and Tollrian (2000).
  13 See also Curry and DeMichele (1977).
                                                       Ecological Monographs
                       JONATHAN M. JESCHKE ET AL.
98
                                                           Vol. 72, No. 1

TABLE 1.  Extended.

                                        Model
                   FF GG HH     II  JJ  KK  LL MM NN OO      PP  QQ RR SS     TT  UU
       Features        (CP) (H) (H)   (C)  (H)  (H)  (H) (H) (H) (C)    (C)  (C) (CH) (H)   (C)  (C)
Components
 Success rate1
 Probability of attack
 Handling time2
 Searching and handling over-
  lapping3
 Hunger and satiation4
 Handling prey digesting prey
 Adaptive behavior5
 Incomplete consumption6
 Nonforaging activities7
 Spatial heterogeneity8
 Temporal heterogeneity9
 Stochasticity10
 Environmental conditions11
 Predator injury by prey
 Inducible defenses12

Dependent on prey density
 Prey density
 Decreasing prey density13
 Alternative prey14
 Learning or switching15
 Swarming effect16

Dependent on predator density
 Predator density
 Interference between predators17
 Multiple predator effects18

Functional response types
  Type I
  Type II
  Type III
  Dome shaped
  Other forms19
  14 See also Belovsky (1978, 1984a, b, c, 1986a, b, 1987), Owen-Smith and Novellie (1982), Engen and Stenseth (1984),

Metz and van Batenburg (1985a), Wanink and Zwarts (1985), Abrams (1987, 1989, 1990b, c), Belovsky et al. (1989), Abrams
and Matsuda (1993), Doucet and Fryxell (1993), Parsons et al. (1994), Forchhammer and Boomsma (1995), Fryxell and
Lundberg (1997), Rothley et al. (1997), Schmitz et al. (1997), and Wilmshurst et al. (2000).
  15 Learning includes training effects; switching means either switching between prey types (in this case, there is a ‘‘ ’’

at ‘‘alternative prey’’) or behavioral switching, e.g., from sitting and waiting to cruising. Only those models that explicitly
consider learning or switching have ‘‘ ’’ here. Optimal foraging models where switching is a simulation result have ‘‘ ’’
here. For an experimental example of the interaction between learning and spatial distribution see Real (1979); see also
Fryxell and Lundberg (1997) and Kaiser (1998).
  16 A swarming effect decreases predation rate with increasing prey density. It can be the result of (1) a better or earlier

detection of the predator by prey, (2) a worse detection of prey by the predator, (3) a better active prey defense, (4) predator
confusion which usually decreases probability or efficiency of attack, (5) clogging of filters (in case of filter feeders), or (6)
accumulation of toxic prey substances. The form of the functional response can be dome shaped in this case. See Miller
(1922), Brock and Riffenburgh (1960), Mori and Chant (1966), Tostowaryk (1972), Halbach and Halbach-Keup (1974), Neill
and Cullen (1974), Nelmes (1974), Milinski and Curio (1975), Bertram (1978), Lazarus (1979), Williamson (1984), Morgan
and Godin (1985), Landeau and Terborgh (1986), and Inman and Krebs (1987).
  17 Interference also includes prey exploitation by other predators. Only those models that consider interference inclusively

and prey exploitation explicitly have ‘‘ ’’ here. Models that account for a decreasing prey density through predation and the
number of predators present and include prey exploitation in an implicit way have ‘‘ ’’ here. See also models by Griffiths and
Holling (1969), Hassell and Varley (1969), Royama (1971, model in §4i), DeAngelis et al. (1975), Curry and DeMichele (1977),
Parker and Sutherland (1986), Korona (1989), Ruxton et al. (1992), Holmgren (1995), Fryxell and Lundberg (1997), and
Doncaster (1999); for empirical studies, see Norris and Johnstone (1998), Triplet et al. (1999), or references in Holmgren (1995).
  18 Soluk (1993), Sih et al. (1998).
  19 Crowley (1973) and Farnsworth and Illius (1996), intermediate type I/II; Nakamura (1974), type II similar; Lam and

Frost (1976), Fujii et al. (1986), type I similar; Lehman (1976), partly type I similar; Metz et al. (1988) and random patch
             ˚¨
model (Lundberg and Astrom 1990), hyperbolic (type II similar) functional response without an asymptote; Abrams (1982),
Juliano (1989), Fryxell (1991), Schmitz (1995), Hirakawa (1997b), Farnsworth and Illius (1998), Abrams and Schmitz (1999),
and Berec (2000), various forms; see also Parsons et al. (1967, type II with a threshold prey density, corresponding empirical
curves in the same study and in Parsons et al. (1969), Cook and Cockrell (1978; double plateau functional responses), Abrams
(1987, 1989; decreasing functional responses), and Fryxell and Lundberg (1997; various forms).
                   PREDATOR FUNCTIONAL RESPONSES
Febraury 2002                                                     99




                 FIG. 1.  A ‘‘family tree’’ of functional response models.




All functional response models include a factor that     Although purely phenomenological, the Gause-Ivlev
determines the curve’s gradient at the origin (‘‘success   equation (Gause 1934, Ivlev 1961) has usually been
rate’’ which is a measure of the predator’s hunting ef-    viewed as the classical satiation model
ficiency; it has been termed the ‘‘rate of successful
                                      y(x)  ymax (1  exp [ a x])      (1)
search’’ by Holling [1959a, b, 1965, 1966]).
                               where a is hunting success (dimension in SI units: m2
        Handling and digestion            for a two-dimensional system, e.g., a terrestrial system,
                               and m3 for a three-dimensional system, e.g., an aquatic
  Limitation of predation rate at high prey densities
                               system), x is prey density (individuals/m2 or individ-
has usually been attributed to either handling time or
                               uals/m3, respectively), y is predation rate (s 1), and ymax
satiation. However, the exact nature of these two factors
                               is asymptotic maximum predation rate as x approaches
and their relationship has been modeled in a variety of
                               infinity (s 1). In the common interpretation, the diges-
different ways and this has led to considerable con-
                               tive system determines ymax, and the functional response
fusion. The point is that handling prey is an active
                               curve gradually rises to this value. Rashevsky (1959)
process whereas digestion is a background process. As
                               has extended the Gause-Ivlev equation by modeling
a consequence, in contrast to handling prey, digestion
                               satiation more mechanistically: the predator’s gut con-
does not directly prevent the predator from further
                               tent is increased by ingestion and decreased by diges-
searching or handling. Rather, digestion influences the
                               tion. Other models including satiation but not handling
predator’s hunger level, which in turn influences the
                               time have been developed by Watt (1959), Parsons et
probability that the predator searches for new prey. It
                               al. (1967), Royama (1971), Nakamura (1974), Lam and
is thus necessary to discriminate digestion from han-
                               Frost (1976), Lehman (1976), Sjoberg (1980), Crisp et
                                                 ¨
dling in a functional response model. In the following,
                               al. (1985), Metz and van Batenburg (1985b), Metz et
we briefly review existing models with respect to their
                               al. (1988), Abrams (1990c), Casas et al. (1993; para-
treatment of these two factors. In our opinion, no com-
pletely satisfying solution to the problem exists to date.  sitoid egg load as analogous to hunger level), Henson
  Models including satiation but not handling time.—     and Hallam (1995), and Abrams and Schmitz (1999).
                                                          Ecological Monographs
                         JONATHAN M. JESCHKE ET AL.
100
                                                              Vol. 72, No. 1


  Models including handling time but not satiation.—         b  tatt /   teat
In contrast, there are a number of models that include
                                      ⇔ asymptotic maximum predation rate
handling time but no predator satiation effects (Holling
1959b, Cushing 1968, Royama 1971, Rogers 1972,                          teat ) 1.
                                          (tatt /                      (4)
Tostowaryk 1972, Beddington 1975, Hassell et al.
                                In using this definition, handling time includes time
1977, Real 1977, Cook and Cockrell 1978, Curry and
                                wasted through unsuccessful attacks (see also Mills
Feldman 1979, Longstaff 1980, McNair 1980, Visser
                                1982, Abrams 1990a, Streams 1994).
and Reinders 1981, Abrams 1982, 1987, 1990a, Fujii
                                  Models including both handling time and satia-
et al. 1986, Dunbrack and Giguere 1987, Ungar and
                 `
                                tion.—One approach to include both handling time and
                         ˚¨
Noy-Meir 1988, Juliano 1989, Lundberg and A strom
                                digestion time is to sum them up or to increase handling
1990, Spalinger and Hobbs 1992, Parsons et al. 1994,
                                time by a ‘‘digestive pause’’ (Crowley 1973, Rao and
Ruxton and Gurney 1994, Streams 1994, Farnsworth
                                Kshirsagar 1978, Mills 1982, Henson and Hallam
and Illius 1996, Fryxell and Lundberg 1997, Cosner et
                                1995), i.e., an inactive time period related to digestion
al. 1999, Berec 2000). The most popular functional
                                (Holling 1965, 1966). When modeled this way, diges-
response model today, Holling’s (1959b) disc equation,
                                tion is not distinguished from handling. Mills (1982)
belongs to this class:
                                used this concept to extend the disc equation by inter-
                     ax           preting its parameter b as
             y(x)               (2)
                      abx
                   1
                                   b  tatt   teat   stdig
where a is success rate (dimension in SI units: m2/s or
                                      ⇔ asymptotic maximum predation rate
m3/s, respectively; note that the dimensions of a and
a [Gause-Ivlev equation] differ), b is predator han-                              1
                                               teat  stdig )
                                          (tatt                       (5)
dling time per prey item (s), x is prey density (indi-
                                where s is satiation per prey item (dimensionless) and
viduals/m2 or individuals/m3, respectively), and y is
                                tdig is digestion time per prey item (s; see Table 2).
predation rate (s 1). The curve’s gradient at the origin
                                  A second way to consider both handling and diges-
is equal to a, and the asymptotic maximum for x as x
                                tion time is to combine the disc equation (which already
approaches infinity is 1/b. The disc equation is math-
                                includes handling time) with a digestive capacity con-
ematically equivalent to the Michaelis-Menten model
                                straint (Fryxell 1991, Schmitz 1995, Hirakawa 1997b,
of enzyme kinetics and the Monod formula for bacterial
                                Farnsworth and Illius 1998). This constraint limits
growth. The Royama-Rogers random predator equation
                                maximum predation rate but does not otherwise affect
(Royama 1971, Rogers 1972) is a modification of the
                                the functional response. These models therefore dis-
disc equation that accounts for a decreasing prey den-
                                criminate between handling and digesting prey. How-
sity in the course of an experiment or between discrete
                                ever, neither the process of digestion, nor the predator
prey generations. In the original paper (Holling 1959b),
                                satiation level are considered. The approach to combine
the parameter b of the disc equation denoted the general
                                handling time with a digestive capacity constraint has
meaning of ‘‘handling time’’ at that time, i.e., the sum
                                its origins in linear programming models (e.g., Belov-
of attacking time tatt (per prey item; including evalu-
                                sky 1978, 1984a, b, c, 1986a, b, 1987, Doucet and
ating, pursuing, and catching time) and eating time teat
                                Fryxell 1993, Forchhammer and Boomsma 1995).
(per prey item):
                                  The only family of models that treats digestion as a
   b  tatt  teat                      background process, which influences foraging activ-
     ⇔ asymptotic maximum predation rate          ities but does not prevent them, is Holling’s (1966)
                                invertebrate model and its extensions (Holling 1965,
                   1
              teat )
         (tatt                  (3)  Holling and Buckingham 1976, Curry and DeMichele
                                1977, Metz and van Batenburg 1985a). In the inver-
with handling time b as it was originally defined by
                                tebrate model, the predation cycle is subdivided into
Holling (1959b).
                                several stages, and each stage depends on predator hun-
  Holling originally developed the disc equation as a
                                ger level. After a meal, the predator is assumed to
mechanistic model for an artificial predator–prey sys-
                                undergo a digestive pause and then continues searching
tem: humans ‘‘preying’’ on paper discs (Holling
                                when it is hungry again. While searching, the predator
1959b). Compared to natural predator–prey systems,
                                simultaneously continues digestion of its last meal. The
however, it is now clear that the underlying assump-
                                invertebrate model therefore discriminates between
tions are unrealistic (Hassell et al. 1976). Two points
                                handling and digesting prey. Here, the length of the
have met the most severe criticism: First, the predator
                                digestive pause depends on hunger level. Since hunger
does not become satiated, and second, the disc equation
                                level in turn depends on prey density, the length of the
assumes that every attack by the predator is successful,
                                digestive pause depends on prey density. This is in
i.e., attack efficiency    100%. Attack efficiencies
  100% can be incorporated into the disc equation by      contrast to the models mentioned above (Eq. 5), where
defining b as                          the length of the digestive pause is unrealistically as-
                    PREDATOR FUNCTIONAL RESPONSES
Febraury 2002                                                       101

TABLE 2.   The SSS equation parameters.

Parameter                  Description                  Dimension†   Defined for
                                                 m3/s
        Encounter rate   number of encounters between a searching predator and             [0;  ]
         a single prey item; an encounter is defined as an arrival of a prey item
         in the predator’s encounter volume
        Probability that the predator detects encountered prey                      [0; 1]
        Efficiency of attack   proportion of successful attacks                     [0; 1]
   s     Satiation per prey item  reciprocal capacity of the hunger-determining             [0; ]
         part of the gut (mostly stomach or crop); example: if the stomach ca-
         pacity of a human is equal to 10 potatoes, then s   0.1
   tatt    Attacking time per prey item   time between prey detection and end of      s       [0;  ]
         attack
   tdig    Digestion time per prey item   food transit time ( 50% emptying time)      s       [0;  ]
         for the hunger-determining part of the gut, e.g., stomach transit time
         for humans
   teat    Eating time per prey item   time between capture and finished ingestion      s       [0;  ]
 Note: The parameters can be summarized by a (success rate [m3/s]), b (corrected handling time [s]), and c (corrected
digestion time [s]), see Eq. 13.
 † In SI units and given for a three-dimensional system, e.g., an aquatic system; in the case of a two-dimensional system,
e.g., a terrestrial system, m3 must be replaced by m2.



sumed to be constant. The term ‘‘digestive pause’’ re-     is no mechanistic linkage to the processes of ingestion
lates to foraging activities only: predators may well use   and digestion.
the digestive pause for nonforaging activities, for ex-      Their mathematical simplicity renders the disc, the
ample, for looking out for top predators or for sleeping.   random predator, and the Gause-Ivlev equation as func-
However, because of its 22 parameters, the invertebrate    tional response submodels in predator–prey population
model is extremely unwieldy, and its extensions are      models. However, for a deeper understanding of the
even more elaborate.                      functional response, mechanistic models are necessary.
                                The parameters of mechanistic models can all be mech-
    Phenomenological vs. mechanistic models         anistically explained. These models can thus, for ex-
                                ample, be used to calculate the effects of changing
  Probably because of their mathematical simplicity,
                                predator or prey characteristics (e.g., defenses) on pre-
the Holling (1959b) disc equation (Eq. 2), the Royama-
                                dation rate.
Rogers random predator equation (Royama 1971, Rog-
ers 1972), and the Gause-Ivlev equation (Gause 1934,
                                 THE STEADY-STATE SATIATION (SSS) EQUATION
Ivlev 1961; Eq. 1) have been the most popular func-
                                 We have shown that a handy mechanistic functional
tional response models. However, they must be con-
                                response model that realistically incorporates handling
sidered phenomenological. That is, although they cor-
                                and digesting prey has been lacking so far. In this sec-
rectly reproduce the shape of natural (type II) func-
                                tion, we therefore develop such a model: the steady-
tional responses, they are not able to explain the un-
                                state satiation (SSS) equation. It is based on the disc
derlying mechanism; or, in other words, its parameters
                                equation and divides the predation cycle into five stag-
cannot all be mechanistically explained. In the cases
                                es: search, encounter, detection, attack, and eating (Fig.
of the disc equation and the random predator equation,
                                2). We assume that these stages are mutually exclusive.
the parameter a (success rate) can be mechanistically
                                Each stage is characterized by two components: The
explained (Holling 1966, Ungar and Noy-Meir 1988,
                                amount of time needed for its completion and the con-
Streams 1994, Hirakawa 1997b; see also Eq. 7 below),
                                ditional probability that the predator reaches this stage
but not the parameter b (handling time). When fitting
                                given that it has reached the previous one (exception:
the disc equation or the random predator equation to
                                encounter; here, it is not a probability but a rate; note
an empirical curve, the resulting value for b is a mixture
                                that the encounter rate can have a value larger than
of different biological processes (Table 1) including
                                unity). Digestion is modeled as a background process
handling (attacking and eating) and digestion. As we
                                influencing the predator’s hunger level, which in turn
have pointed out above, handling is an active process,
                                determines the probability that the predator searches
whereas digestion is a background process. They can-
                                for prey.
not be adequately condensed into only one parameter.
In the case of the Gause-Ivlev equation, neither param-
                                   The SSS equation components, parameters,
eter can be mechanistically explained. Its parameter a
                                         and assumptions
(hunting success) differs in its dimension from the pa-
rameter a of the disc equation and the random predator      The SSS equation components are given in Table 1
equation; a lacks a mechanistic explanation. The other     and its parameters in Table 2. Like nearly every model,
parameter, ymax, is just the asymptote of the curve; there   the SSS equation is a compromise between realism and
                                                     Ecological Monographs
                       JONATHAN M. JESCHKE ET AL.
102
                                                         Vol. 72, No. 1


                                   1) There is only a single predator and a single type
                                 of prey.
                                   2) The prey density is constant.
                                   3) Prey are independently and randomly distributed.
                                   4) Stages of the predation cycle exclude each other
                                 (Fig. 2).
                                   5) The probability that the predator searches (under
                                 the condition that it is not handling prey), (x), is di-
                                 rectly proportional to the predator’s hunger level h(x).
                                   6) The hunger level h(x) depends on the fullness of
                                 a certain part of the gut (e.g., stomach, crop).
                                   7) The hunger level h(x) at a given prey density x is
                                 in a steady state, which is determined by an equilibrium
                                 of ingestion and digestion.
                                   8) The probability of attack, , is unity, i.e., when-
                                 ever a searching predator encounters and detects a prey,
                                 it will attack.
                                   9) The probability that the predator detects encoun-
                                 tered prey, , the efficiency of attack (i.e., the pro-
                                 portion of successful attacks), the attacking time tatt (per
                                 prey item), the eating time teat (per prey item), and the
                                 digestion time tdig (per prey item) are constant.
                                           The SSS equation
                                   To develop the SSS equation, we start with the disc
                                 equation and modify it sequentially. In step 1, each
                                 stage of the predation cycle is included explicitly; in
                                 step 2, predator satiation is included by influencing the
  FIG. 2. The predation cycle. We divide the predation cycle
into five stages: search, encounter, detection, attack, and eat-  probability of searching.
ing. A predator enters a predation cycle under the probability
                                   The stages of the predation cycle are (1) search, (2)
to search, (x); this is determined by the predator’s hunger
                                 encounter, (3) detection, (4) attack, and (5) eating (Fig.
level, which in turn is influenced by digestion time. Then the
                                 2). The probabilities that a predator reaches these stages
predator successively reaches the following stages. The prob-
ability that the predator reaches a stage under the condition   are (1) the probability that a predator not occupied
that it has reached the previous stage is given in the corre-   with handling searches for prey, (2) the encounter rate
sponding arrow, e.g., the probability that the predator detects
                                   between a searching predator and an individual prey,
a prey under the condition that it has encountered that prey
                                 (3) the probability that the predator detects an en-
is (exception: is not a probability but a rate [encounter
                                 countered prey individual, (4) the probability that the
rate]; note that it can be larger than unity). Since is set as
unity in the SSS equation (assumption 8), it is given in pa-   predator attacks a detected prey individual, and (5) the
rentheses here. Terms in circles indicate time demands of
                                 probability that an attack is successful, i.e., the ef-
corresponding stages per prey item. We assume that the stages
                                 ficiency of attack. We now incorporate these probabil-
are mutually exclusive (assumption 4). Terms with a super-
                                 ities into the disc equation.
script ‘‘a’’ determine predator success rate (a).
                                   The searching probability (x).—In the disc equa-
                                 tion, the predator shows only two kinds of behavior:
                                 searching for and handling prey. Therefore, the prob-
applicability. It is more realistic than the disc equation,
                                 ability that the predator searches for prey under the
but it is reductive compared to nature. The SSS equa-
                                 condition that it is not handling prey, (x), is unity. To
tion is a conceptual model that can, for example, be
                                 allow values below unity, (x) has to be incorporated
used to assess how changing predator or prey charac-
                                 explicitly into the disc equation:
teristics (e.g., defenses) qualitatively affect the func-
tional response. The point of the SSS equation is not                      (x)ax
                                          y(x)         .         (6)
to quantitatively predict real functional response
                                              1   (x)abx
curves. It is therefore not necessary to incorporate too
                                        depends on prey density x because it is
many features into the model, which would render it        Note that
unwieldy. However, extensions for specific predator–        affected by the predator’s hunger level (see the next
prey systems are possible; these will allow us to make      paragraph and assumption 5), which in turn depends
quantitative predictions with the model as well. For       on prey density (see the next paragraph and assumption
this purpose, references given in Table 1 may be help-      7): (h)   (h(x))   (x).
                                   The encounter rate , the probability of detection ,
ful.
                                 the probability of attack , and the efficiency of attack
  The assumptions of the model are as follows:
                          PREDATOR FUNCTIONAL RESPONSES
Febraury 2002                                                                     103


                                                        c · y(x))ax
.—The product of all these terms is predator success                           (1
                                            y(x)                  .           (12)
rate a. However, for simplicity, we set    1 (as-                               c · y(x))abx
                                                 1     (1
sumption 8). Thus,
                                  Solving for y(x) finally gives the following SSS equation:
             a        .          (7)
                                      1
The encounter rate    can be calculated by various




                                      
                                                                           c) 2 )
                                           ax(b    c)        ax(2(b       c)  ax(b
                                                      1
formulae from different authors. For a three-dimen-
                                                         2abcx
sional model, e.g., in aquatic systems, one may use
the equation given by Gerritsen and Strickler (1977).                         a, b, c, x     0
For an analogous two-dimensional model, e.g., in ter-
                                           ax
restrial systems, see Koopman (1956), and for a three-                      b            c
                                                      0           0
                                      1
                                  y(x)




                                      
                                            abx
dimensional model with a cylindrical instead of a
                                           ax
spherical encounter volume, see Gigue re et al. (1982).
                    `
                                                 b            c
                                                      0           0
For further models, see Royama (1971: Eq. 4e.6), Get-                 acx
                                        1
ty and Pulliam (1991), Parsons et al. (1994), Hirakawa
                                        ax        b     c     0
(1997b), and reviews from Schoener (1971) and Curio
                                      0          a            x
                                                      0   or       0
(1976). Here, for simplicity,   is not calculated by
one of these formulae but is a model input; probability
                                                                           (13)
of detection and efficiency of attack are also model
                                  with success rate a       , corrected handling time
inputs.
                                  b t att /  t eat , and corrected digestion time c st dig .
  Explicitly incorporating efficiency of attack allows
                                    For details on deriving Eq. 13 from Eq. 12, see Ap-
us to account for time wasted through unsuccessful
                                  pendix A. For c 0 (i.e., no satiation), the SSS equa-
attacks. Thus, handling time b can be calculated ac-
                                  tion simplifies to the disc equation but with the defi-
cording to Eq. 4.
                                  nitions of Eq. 13 for a and b. For b      0 (i.e., zero
  The second and final step in deriving the SSS equa-
                                  handling time), the SSS equation simplifies to the disc
tion is to incorporate digestion. We do this by assuming
                                  equation but with c instead of b, i.e., digestion time
that
                                  replaces handling time in this case. Finally, without
                   h(x).
             (x)                 (8)
                                                       c
                                  any handling time or satiation (b       0), there are
                                  no density dependent effects and so, predation rate is
This is assumption 5 and is also assumed by Rash-
                                  directly proportional to prey density.
evsky (1959). The hunger level h(x) is the proportion
of empty volume of that part of the gut that is re-
                                           Properties of the SSS equation
sponsible for feelings of hunger and satiation in the
                                    The SSS equation produces type II functional re-
predator under consideration (mostly stomach or
                                  sponses (Fig. 3). As in the disc equation, the gradient
crop); h(x) is defined for [0; 1], where h  0 means
                                  at the origin is equal to the predator’s success rate a:
no hunger, i.e., full gut, and h 1 means 100% hunger,
i.e., empty gut. Empirical studies usually find a hy-
                                                      dy(x)
perbolic relationship between starvation time and                                     a.
                                                 lim                        (14)
                                                      dx
                                                 x→0
hunger level, e.g., Holling (1966) for mantids (Hier-
odula crassa and Mantis religiosa), Antezana et al.         The asymptotic maximum predation rate for prey den-
(1982) for krill (Euphausia superba), Hansen et al.         sity as x approaches infinity is
(1990) for copepods (Calanus finmarchicus), and sev-
                                                             c) 2
                                               b    c     (b            1
eral works on fish (reviewed by Elliott and Persson             lim y(x)                               .
                                                                    max(b; c)
                                                      2bc
1978). This hyperbolic relationship can be described            x→

by the following differential equation:
                                  where, for handling-limited predators,
        dh(x)         h(x)
             1
                                                                 1
                       sy(x).     (9)
                                                   c ⇔ lim y(x)
                                               b
         dt        tdig
                                                                 b
                                                       x→
Since we assume a constant prey density (assumption
                                  and, for digestion-limited predators,
2), the equilibrium hunger level can be obtained by
setting dh(x)/dt  0, giving                                                   1
                                                 b ⇔ lim y(x)
                                               c                   .         (15)
                                                                 c
                               (10)
         h(x)        s·tdig·y(x).
               1                                       x→

                                  Thus, the larger one of the two terms b and c determines
We define c s·tdig as ‘‘corrected digestion time’’, i.e.,
                                  the asymptotic maximum predation rate. This is, be-
digestion time corrected for gut capacity. Therefore,
                                  cause digestion is a ‘‘background process’’, i.e., han-
          h(x)         c·y(x).
                 1             (11)  dling and digestion can be carried out simultaneously.
                                  The slower one of these two processes is then limiting.
Inserting Eq. 11 into Eqs. 8 and 6 yields
                                                       Ecological Monographs
                       JONATHAN M. JESCHKE ET AL.
104
                                                           Vol. 72, No. 1


                                    When corrected handling time exceeds corrected di-
                                  gestion time (b c, condition 1), the asymptotic max-
                                  imum predation rate is 1/b. This is the same situation
                                  as in a disc equation when attack efficiencies 100%
                                  are considered (see Eq. 4). We call predators under this
                                  condition ‘‘handling-limited predators.’’ Fig. 3a shows
                                  graphs of the SSS equation for handling-limited pred-
                                  ators. Although the asymptote is independent of c, it
                                  is approached more slowly as digestion time becomes
                                  more important, i.e., large digestion times result in a
                                  slower rise of the curve. As c approaches 0, the SSS
                                  curve approaches a disc equation curve (with a cor-
                                  rection for attack efficiencies 100%).
                                    When corrected digestion time exceeds corrected
                                  handling time (c b, condition 2), the asymptotic max-
                                  imum predation rate equals 1/c. We call predators under
                                  this condition ‘‘digestion-limited predators.’’ Fig. 3b
                                  shows graphs of the SSS equation for digestion-limited
                                  predators. With larger handling times, the asymptote is
                                  approached more slowly, yet the asymptote itself is
                                  independent of b. As b approaches 0, the SSS curve
                                  approaches a disc equation curve with digestion in
                                  place of handling (c instead of b) as the limiting factor.
                                    SSS equation curves are more flexible than disc
                                  equation curves. Thus, it is impossible to satisfyingly
                                  fit the disc equation to a SSS equation curve (with the
                                  exceptions b   0 or c  0). This is, because, in the
                                  disc equation, one parameter (b) determines the curve’s
                                  asymptote, and two parameters (a and b) determine
                                  how the curve reaches this asymptote, i.e., the curve’s
                                  slope. In contrast, in the SSS equation, one parameter
                                  (the larger one of the parameters b and c) determines
                                  the curve’s asymptote, and three parameters (a, b, and
                                  c) determine how the curve reaches this asymptote.
                                    Fig. 3c illustrates how time wasted through unsuc-
                                  cessful attacks (attack efficiency    100%) reduces
                                  the slope of the functional response curve (and, in case
                                  of handling-limited predators, the asymptotic maxi-
                                  mum predation rate).

                                              DISCUSSION
                                   We have developed a handy mechanistic functional
                                  response model (the SSS equation) that realistically

  FIG. 3. Graphs of the SSS equation (Eq. 13). (A) Han-
dling-limited predators. Model inputs were success rate a      ←
2, corrected handling time b 0.02, and corrected digestion
time c 0, 0.01, or 0.02, respectively; thus, b c. All curves    (C) Effect of attack efficiency . Model inputs were a     2
                                                     0.25), b 0.01 (
are type II functional responses, and for all curves, asymptotic  (corresponding    0.5) or 1 (             0.5,
                                  tatt 0.0025, teat 0.005; b [Eq. 13] 0.0025 / 0.5 0.005
maximum predation rate 1/b 50 (Eq. 15). However, this
                                              0.25, tatt 0.0025, teat 0.005; b [Eq.
asymptotic maximum is approached more slowly as digestion        0.01) or 0.015 (
time becomes more important. For c                                   0.015), and c
                      0, the SSS equation   13]   0.0025 / 0.25  0.005          0.02. When
is equal to the disc equation (Eq. 2). (B) Digestion-limited    attack efficiency is halved (from 0.5 to 0.25), the gradient at
predators. Model inputs were a     2, b   0, 0.01, or 0.02,  the origin is halved (a   2 or 1, Eq. 14) and the predation
respectively, and c   0.02; thus, c  b. All curves are type  rate is decreased at almost all prey densities. However, in
II functional responses, and for all curves, asymptotic max-    case of a digestion-limited predator (as in our example), as-
imum predation rate     1/c  50 (Eq. 15). However, this    ymptotic maximum predation rate remains constant (1/c
asymptotic maximum is approached more slowly as handling      50, Eq. 15). In the case of a handling-limited predator (graph
time becomes more important. For b                 not shown), b is increased, and thus asymptotic maximum
                      0, the SSS equation
is equal to the disc equation, when b is replaced by c there.    predation rate is decreased.
                   PREDATOR FUNCTIONAL RESPONSES
Febraury 2002                                                     105


incorporates success rate, handling time, and satiation.   of top predators, migration, molting, reproductive ac-
The satiation level is assumed to linearly decrease hunt-  tivities, resting, sleeping, territorial behavior, thermo-
ing activities. The SSS equation thereby fills a gap in    regulation, or times of slow rates of metabolism like
functional response theory, because previous models     winter dormancy).
either do not treat satiation in a realistic way (since     The easiest method to detect a handling-limited pred-
they do not discriminate between handling and di-      ator is to directly measure corrected handling time (ac-
gesting prey or simply include satiation by a maximum    cording to Eq. 4) as well as corrected digestion time
predation rate, i.e., a digestive capacity constraint) or  (according to Eq. 13) and to compare them. However,
are extremely unwieldy.                   all predators, from whom both measurements are avail-
  Like the widely used disc equation, the SSS equation   able in the literature, are digestion-limited (see next
produces type II functional response curves. However,    section).
there are several differences. First, because of its third   Another method to detect a handling-limited pred-
parameter, the SSS equation is more flexible than the     ator is:
disc equation. The differences are largest when han-      1A) Through observation, directly measure predator
                               handling time b according to Eq. 4.
dling time and digestion time are of the same order of
magnitude (Fig. 3). On the contrary, if one of these      1B) (Alternative to 1A) Perform short-term feeding
two factors is negligibly small, the curve becomes vir-   experiments to get a short-term functional response
tually identical to that of the disc equation. Second, the  without satiation effects. Fit the disc equation (if eaten
disc equation assumes an attack efficiency equal to      prey was replaced) or the random predator equation (if
                               eaten prey was not replaced) to the data to get b (han-
100%. When this is not the case, the maximum pre-
dation rate is decreased because of time spent for un-    dling time according to Eq. 4).
                                2A) Measure long-term maximum feeding rate ymax
successful attacks. Although mentioned by Mills
(1982), Abrams (1990a), and Streams (1994), this ef-     (with satiation) at an extremely high prey density.
fect has not been incorporated into most models. It is     2B) (Alternatively to 2A) Perform long-term feeding
contrary to the basic idea of the disc equation that the   experiments, ideally starting with predators in a steady
parameters a and b are independent (Holling 1965,      hunger state, or do a field study. Fit the disc equation
                               or the random predator equation to the data to get ymax.
1966). In nature, predator attack efficiencies seldom
                                3) If b
reach 100% (see Curio 1976, Vermeij 1982, and Packer            1/ymax, it is likely that the predator is han-
and Ruttan 1988). Taking unsuccessful attacks into ac-    dling limited.
count is especially important for predators with non-      We have applied this method to available literature
negligible attacking times. Third and most important,    data and have found three candidates for handling-lim-
the disc equation (with b interpreted as in Eq. 5) does   ited predators. First, in the host–parasitoid system Silo
                               pallipes (Trichoptera: Goeridae)–Agriotypus armatus
not discriminate between handling and digesting prey.
The SSS equation, on the other hand, takes into account   (Hymenoptera: Agriotypidae), Elliott (1983) directly
                               measured the handling time of A. armatus and found
their different nature, and as a result, the maximum
                               b 20.0 min. In addition, he fitted the random predator
predation rate (prey density approaches infinity) is not
                                                  19.4–20.1 min, thus b
determined by the sum of time spent for handling and     equation to field data: 1/ymax
                                 1/ymax. Second, in the predator–prey system Och-
digesting prey (as in Eq. 5), but by the maximum of
                               romonas sp. (a heterotrophic flagellate)–Pseudomonas
these two terms. Accordingly, we have classified pred-
ators into handling-limited and digestion-limited pred-   sp. (a bacterium), Fenchel (1982a) directly measured
                               the handling time of Ochromonas as b
ators. Note that this classification only refers to high                            20 s. In ad-
prey densities. At intermediate prey densities, our mod-   dition, he performed long-term experiments (Fenchel
                                         19 s, thus b
el shows that also handling-limited (digestion-limited)   1982b): 1/ymax             1/ymax. Third, in the
                               predator–prey system Polinices duplicatus (a naticid
predators experience diminished feeding rates because
of time spent for digesting (handling) prey (Figs. 3a,    gastropod that drills through the shells of its prey)–
                               Mya arenaria (Bivalvia), the handling time of P. du-
b).
                               plicatus in the long-term enclosure experiments of Ed-
        Handling-limited predators          wards and Huebner (1977) can be estimated by data
  Handling-limited predators handle (corrected for at-   from Edwards and Huebner (1977) and Kitchell et al.
                               (1981; Appendix B): b      1.4 d; 1/ymax 1.6 d, thus b
tack efficiencies 100%) prey slower than they digest
them. For parasites and parasitoids, this means that they    1/ymax. Similarly, Boggs et al.’s (1984) results have
                               indicated that P. duplicatus is also handling limited
handle hosts slower than they produce eggs. In han-
                               when feeding on another bivalve, Mercenaria mercen-
dling-limited predators, therefore, prey uptake increas-
                               aria. In their study, P. duplicatus spends 75% of its
es with the amount of time spent for searching and
                               time in handling (i.e., drilling and eating) M. mercen-
handling prey. We consequently expect that, indepen-
                               aria; total foraging time (i.e., searching time plus han-
dent of prey density, handling-limited predators forage
almost all of their available time (i.e., the time not    dling time) was therefore at least 75%. This exceeds
needed for nonforaging activities, such as avoidance     by far corresponding values for digestion-limited pred-
                                                 Ecological Monographs
                    JONATHAN M. JESCHKE ET AL.
106
                                                     Vol. 72, No. 1


ators (see Discussion: Digestion-limited predators),    moschatus], Forchhammer and Boomsma [1995];
                              sheep [Ovis aries], Blaxter et al. [1961]; shrews [Sorex
corroborating our expectation that handling-limited
                              araneus, S. caecutiens, S. isodon], Saariko and Hanski
predators should spend more time in foraging than di-
gestion-limited predators.                 [1990]).
  Further examples for handling-limited predators can    For digestion-limited predators, the SSS equation,
likely be found in other parasitoids, protozoans, and   contrary to Holling’s (1959b) disc equation (Eq. 2),
drilling gastropods. In general, however, handling-lim-  predicts that foraging time decreases with increasing
ited predators seem to be rare.              prey density. This is in accordance with empirical data,
                              for example from birds (Spotted Sandpipers [Actitis
       Digestion-limited predators          macularia], Maxson and Oring [1980]; Verdins [Au-
                              riparus flaviceps], Austin [1978]; Oystercatchers [Hae-
  Digestion-limited predators digest prey items slower
                              matopus ostralegus], Drinnan [1957]; Yellow-eyed
than they handle them. For parasites and parasitoids,
                              Juncos [Junco phaeonotus], Caraco [1979]; humming-
this means they produce eggs slower than they handle
                              birds [Selasphorus rufus], Hixon et al. [1983]) and
hosts. At high prey densities, therefore, predation up-
                              mammals (horses [Equus caballus], Duncan [1980];
take does not further increase with the amount of time
                              white-tailed jackrabbits [Lepus townsendii], Rogowitz
spent for searching and handling prey. This releases
                              [1997]; sheep [Ovis aries], Alden and Whittaker
trade-off situations at high prey densities and closes
                              [1970]; mouflon [Ovis musimon], Moncorps et al.
the gap between optimal foraging and satisficing theory
(J. Jeschke, personal observation; for satisficing, see   [1997]; reindeer [Rangifer tarandus tarandus], Trudell
                              and White [1981]; greater kudus [Tragelaphus strep-
Herbers [1981] and Ward [1992, 1993]).
                              siceros], Owen-Smith [1994]).
  The vast majority of predators seems to be digestion
limited (see also Weiner 1992). Examples have been      Finally, natural predators generally spend a major
                              part of their time in resting. For example, Amoeba pro-
reported from mollusks (veliger larvae: Crisp et al.
[1985]; common or blue mussel [Mytilus edulis], Bayne   teus, Woodruffia metabolica, African Fish Eagles (Hal-
et al. [1989]), crustaceans (Branchipus schaefferi,    iaetus vocifer), lions (Panthera leo), and wild dogs
Streptocephalus torvicornis, Dierckens et al. [1997];   (Lyaon pictus) spend only 17% of their time in hunt-
Calliopius laeviusculus, DeBlois and Leggett [1991];    ing and eating (reviewed by Curio 1976). For further
Daphnia spp., Rigler [1961], McMahon and Rigler      examples, see Herbers (1981) or Bunnell and Harestad
[1963], Geller [1975]; Calanus pacificus, Frost [1972];   (1990). Since resting may be caused by satiation, this
other copepods: Paffenhofer et al. [1982], Christoffer-
             ¨                 may suggest that such predators are digestion limited.
sen and Jespersen [1986], Head [1986], Jonsson and     It is, however, more reliable, to compare predator for-
Tiselius [1990]), insects (Chaoborus spp. larvae, re-   aging and nonforaging times with actual measurements
viewed by Jeschke and Tollrian [2000]; the grasshop-    of handling and digestion time. This approach reveals
pers Circotettix undulatus, Dissosteira carolina, Me-   that the time various herbivores spend for feeding can
lanoplus femur-rubrum, and Melanoplus sanguinipes,     usually be predicted solely from their handling and
Belovsky [1986b]; dusty wing larvae [Conwentzia hag-    digestion times (J. Jeschke, personal observation). In
eni], green lacewing larvae [Chrysopa californica], red  other words, resting often seems to be motivated by
mite destroyer larvae [Stethorus picipes], Fleschner    satiation.
[1950]), birds (Woodpigeons [Columba palumbus],
                                   Applications of the SSS equation
Kenward and Sibly [1977]; Oystercatchers [Haema-
topus ostralegus], Kersten and Visser [1996]; hum-      The SSS equation was designed as a conceptual mod-
mingbirds [Selasphorus rufus], Hixon et al. [1983], Di-  el for developing general and qualitative predictions
amond et al. [1986]), and mammals (moose [Alces al-    about functional responses. It can be used to predict
ces], Belovsky [1978]; pronghorn antelopes [Antilo-    the effects of changing predator or prey characteristics
capra americana], bison [Bison bison], elk [Cervus     by analyzing changes of the corresponding parameters.
elaphus], yellow-bellied marmots [Marmota flaviven-     For example, the effects of different kinds of prey de-
tris], mule deer [Odocoileus hemionus], white-tailed    fenses can be predicted. A defense that reduces the
deer [Odocoileus virginianus], bighorn sheep [Ovis     predator’s success rate (e.g., camouflage) will have its
canadensis], Columbian ground squirrels [Spermophi-    largest effects at low prey densities. In contrast, an
lus columbianus], Rocky Mountain cotton tails [Syl-    increase in handling time due to a defense (e.g., an
vilagus nuttali], Belovsky [1986b]; cattle [Bos taurus],  escape reaction [decreases success rate and increases
Campling et al. [1961]; beavers [Castor canadensis],    handling time]) will lower maximum predation rate in
Belovsky [1984b], Doucet and Fryxell [1993], Fryxell    handling-limited predators. In digestion-limited pred-
et al. [1994]; Thomson’s gazelles [Gazella thomsoni],   ators, either predation rates will decrease or total for-
Wilmshurst et al. [1999]; human beings [Homo sapi-     aging time will increase. Finally, an increase in diges-
ens], Belovsky [1987]; snowshoe hares [Lepus amer-     tion time (e.g., due to barely digestible substances) will
icanus], Belovsky [1984c]; meadow voles [Microtus     lower predation rates at high prey densities in diges-
pennsylvanicus], Belovsky [1984a]; muskoxen [Ovibus    tion-limited predators (see also Jeschke and Tollrian
                    PREDATOR FUNCTIONAL RESPONSES
Febraury 2002                                                       107

                                 result of adaptive consumer behavior. Evolutionary Ecol-
2000). The same considerations can be used to for-
                                 ogy 3:95–114.
mulate hypotheses about optimal investment of pred-
                                Abrams, P. A. 1990a. The effects of adaptive behavior on
ators in raising success rate, handling efficiency, or       the type-2 functional response. Ecology 71:877–885.
digestive capacity. More generally, the SSS equation      Abrams, P. A. 1990b. Mixed responses to resource densities
can be linked with cost–benefit models to investigate       and their implications for character displacement. Evolu-
                                 tionary Ecology 4:93–102.
predator and prey evolution using predation rate as an
                                Abrams, P. A. 1990c. Adaptive responses of generalist her-
indirect measure of fitness.
                                 bivores to competition: Convergence or divergence. Evo-
  Since the basic SSS equation contains many sim-        lutionary Ecology 4:103–114.
plifying assumptions, it should not primarily be viewed     Abrams, P. A. 1991. Life history and the relationship between
                                 food availability and foraging effort. Ecology 72:1242–
as a model for quantitatively predicting functional re-
                                 1252.
sponses. However, the model is open to modifications
                                Abrams, P. A. 1992. Predators that benefit prey and prey that
to better match the properties of specific predator–prey      harm predators: unusual effects of interacting foraging ad-
systems (using numerical analyses when necessary).        aptations. American Naturalist 140:573–600.
For example, making attack efficiency a decreasing        Abrams, P. A. 1993. Why predation rates should not be pro-
                                 portional to predator density. Ecology 74:726–733.
function of prey density allows the modeling of a
                                Abrams, P. A., and H. Matsuda. 1993. Effects of adaptive
swarming effect due to predator confusion. After in-
                                 predatory and anti-predator behavior in a two prey–one
corporating this confusion effect and accounting for a      predator system. Evolutionary Ecology 7:312–326.
decreasing prey density, the model adequately predicts     Abrams, P. A., and O. J. Schmitz. 1999. The effect of risk
                                 of mortality on the foraging behaviour of animals faced
the functional response of Chaoborus obscuripes lar-
                                 with time and digestive capacity constraints. Evolutionary
vae (Diptera) feeding on Daphnia pulex (Crustacea; J.
                                 Ecology Research 1:285–301.
Jeschke, personal observation).                 Alden, W. G., and I. A. M. Whittaker. 1970. The determinants
  Finally, the concept underlying the SSS equation        of herbage intake by grazing sheep: the interrelationship
may be used to improve predator–prey theory in gen-        of factors influencing herbage intake and availability. Aus-
                                 tralian Journal of Agricultural Research 21:755–766.
eral, e.g., optimal foraging theory. Classical optimal
                                Anholt, B. R., D. Ludwig, and J. B. Rasmussen. 1987. Op-
foraging theory is based on the disc equation, consid-
                                 timal pursuit times: how long should predators pursue their
ering handling time but not digestion time (reviewed       prey? Theoretical Population Biology 31:453–464.
by Stephens and Krebs 1986). In models developed        Antezana, T., K. Ray, and C. Melo. 1982. Trophic behavior
                                 of Euphausia superba Dana in laboratory conditions. Polar
primarily for herbivores, handling time is often com-
                                 Biology 1:77–82.
bined with a digestive capacity constraint: ‘‘linear pro-
                                Arditi, R., and L. R. Ginzburg. 1989. Coupling in predator–
gramming models’’ (e.g., Belovsky 1978, 1984a, b, c,       prey dynamics: ratio dependence. Journal of Theoretical
1986a, b, 1987, Doucet and Fryxell 1993, Forchham-        Biology 139:311–326.
                                ˚¨
mer and Boomsma 1995), ‘‘digestive rate models’’        Astrom, M., P. Lundberg, and K. Danell. 1990. Partial con-
                                 sumption by browsers: trees as patches. Journal of Animal
(Verlinden and Wiley 1989, Hirakawa 1997a, b, Farns-
                                 Ecology 59:287–300.
worth and Illius 1998), and patch selection models
                                Austin, G. T. 1978. Daily time budget of the postnesting
(Fryxell 1991, Wilmshurst et al. 1995, 2000). The SSS       verdin. Auk 95:247–251.
equation offers an alternative approach for combining      Bayne, B. L., A. J. S. Hawkins, E. Navarro, and I. P. Iglesias.
handling time with digestion time. Comparing an op-        1989. Effects of seston concentration on feeding, digestion
                                 and growth in the mussel Mytilus edulis. Marine Ecology
timal foraging model based on the SSS equation with
                                 Progress Series 55:47–54.
existing approaches may improve optimal foraging the-
                                Beddington, J. R. 1975. Mutual interference between para-
ory, since one of its major problems has been the lack      sites or predators and its effect on searching efficiency.
of alternative models (e.g., Ward 1992, 1993). This may      Journal of Animal Ecology 44:331–340.
                                Begon, M., J. L. Harper, and C. R. Townsend. 1996. Ecology:
reveal new insights into predator foraging behavior.
                                 individuals, populations, and communities. Third edition.
                                 Blackwell, Oxford, UK.
          ACKNOWLEDGMENTS
                                Belovsky, G. E. 1978. Diet optimization in a generalist her-
 Wilfried Gabriel commented on earlier versions of the SSS    bivore: the moose. Theoretical Population Biology 14:105–
equation and of the manuscript. John Fryxell, Steven Juliano,   134.
Oliver Lange, Herwig Stibor, Matthijs Vos, and two anony-    Belovsky, G. E. 1984a. Herbivore optimal foraging: a com-
mous reviewers commented on the manuscript. Tim Vines       parative test of three models. American Naturalist 124:97–
improved the language. The study was funded by the        115.
Deutsche Forschungsgemeinschaft (DFG).             Belovsky, G. E. 1984b. Summer diet optimization by beaver.
                                 American Midland Naturalist 111:209–222.
          LITERATURE CITED              Belovsky, G. E. 1984c. Snowshoe hare optimal foraging and
                                 its implications for population dynamics. Theoretical Pop-
Abrams, P. A. 1982. Functional responses of optimal for-
                                 ulation Biology 25:235–264.
agers. American Naturalist 120:382–390.
                                Belovsky, G. E. 1986a. Generalist herbivore foraging and its
Abrams, P. A. 1984. Foraging time optimization and inter-
                                 role in competitive interactions. American Zoologist 26:
actions in food webs. American Naturalist 124:80–96.
                                 51–70.
Abrams, P. A. 1987. The functional responses of adaptive
                                Belovsky, G. E. 1986b. Optimal foraging and community
consumers of two resources. Theoretical Population Biol-
                                 structure: implications for a guild of generalist grassland
ogy 32:262–288.
Abrams, P. A. 1989. Decreasing functional responses as a     herbivores. Oecologia 70:35–52.
                                                     Ecological Monographs
                      JONATHAN M. JESCHKE ET AL.
108
                                                         Vol. 72, No. 1

Belovsky, G. E. 1987. Hunter–gatherer foraging: a linear    Curio, E. 1976. The ethology of predation. Springer, Berlin,
                                 Germany.
 programming approach. Journal of Anthropological Ar-
                                Curry, G. L., and D. W. DeMichele. 1977. Stochastic analysis
 chaeology 6:29–76.
                                 for the description and synthesis of predator–prey systems.
Belovsky, G. E., M. E. Ritchie, and J. Moorehead. 1989.
                                 Canadian Entomologist 109:1167–1174.
 Foraging in complex environments: when prey availability
                                Curry, G. L., and R. M. Feldman. 1979. Stochastic predation
 varies over space and time. Theoretical Population Biology
                                 model with depletion. Canadian Entomologist 111:465–
 36:144–160.
                                 470.
Berec, L. 2000. Mixed encounters, limited perception and
                                Cushing, D. H. 1968. Grazing by herbivorous copepods in
 optimal foraging. Bulletin of Mathematical Biology 62:
                                 the sea. Journal du Conseil 32:70–82.
 849–868.
                                DeAngelis, D. L., R. A. Goldstein, and R. V. O’Neill. 1975.
Bernays, E. A., and S. J. Simpson. 1982. Control of food
                                 A model for trophic interaction. Ecology 56:881–892.
 intake. Advances in Insect Physiology 16:59–118.
                                DeBlois, E. M., and W. C. Leggett. 1991. Functional response
Berryman, A. A., A. P. Gutierrez, and R. Arditi. 1995. Cred-
                                 and potential impact of invertebrate predators on benthic
 ible, parsimonious and useful predator–prey models—a re-
                                 fish eggs: analysis of the Calliopius laeviusculus–capelin
 ply to Abrams, Gleeson, and Sarnelle. Ecology 76:1980–
                                 (Mallotus villosus) predator–prey system. Marine Ecology
 1985.
                                 Progress Series 69:205–216.
Bertram, B. C. R. 1978. Living in groups: predators and prey.
                                Demment, M. W., and G. B. Greenwood. 1988. Forage in-
 Pages 64–96 in J. R. Krebs and N. B. Davies, editors.
                                 gestion: effects of sward characteristics and body size.
 Behavioural ecology: an evolutionary approach. Blackwell,
                                 Journal of Animal Science 66:2380–2392.
 Oxford, UK.
                                Diamond, J. M., W. H. Karasov, D. Phan, and F. L. Carpenter.
Blaine, T. W., and D. L. DeAngelis. 1997. The interaction of
                                 1986. Digestive physiology is a determinant of foraging
 spatial scale and predator–prey functional response. Eco-
                                 bout frequency in hummingbirds. Nature 320:62–63.
 logical Modelling 95:319–328.
                                Dierckens, K. R., L. Beladjal, J. Vandenberghe, J. Swings,
Blaxter, K. L., F. W. Wainman, and R. S. Wilson. 1961. The
                                 and J. Mertens. 1997. Filter-feeding shrimps (Anostraca)
 regulation of food intake by sheep. Animal Production 3:
                                 grazing on bacteria. Journal of Crustacean Biology 17:264–
 51–61.
                                 268.
Boggs, C. H., J. A. Rice, J. A. Kitchell, and J. F. Kitchell.
                                Doncaster, C. P. 1999. A useful phenomenological difference
 1984. Predation at a snail’s pace: what’s time to a gastro-
                                 between exploitation and interference in the distribution of
 pod? Oecologia 62:13–17.
                                 ideal free predators. Journal of Animal Ecology 68:836–
Brock, V. E., and R. H. Riffenburgh. 1960. Fish schooling:
                                 838.
 a possible factor in reducing predation. Journal du Conseil
                                Doucet, C. M., and J. M. Fryxell. 1993. The effect of nu-
 25:307–317.
                                 tritional quality on forage preference by beavers. Oikos 67:
Buckner, C. H. 1964. Metabolism, food capacity, and feeding
                                 201–208.
 behavior in four species of shrews. Canadian Journal of
                                Drinnan, R. E. 1957. The winter feeding of the oystercatcher
 Zoology 42:259–279.
                                 (Haematopus ostralegus) on the edible cockle. Journal of
Bunnell, F. L., and A. S. Harestad. 1990. Activity budgets
                                 Animal Ecology 26:439–469.
 and body weight in mammals. How sloppy can mammals
                                Dunbrack, R. L., and L. A. Giguere. 1987. Adaptive re-
                                                   `
 be? Current Mammalogy 2:245–305.
                                 sponses to accelerating costs of movement: a bioenergetic
Campling, R. C., M. Freer, and C. C. Balch. 1961. Factors
                                 basis for the type III functional response. American Nat-
 affecting the voluntary intake of foods by cows. 2. The
                                 uralist 130:147–160.
 relationship between the voluntary intake of roughages, the
                                Duncan, P. 1980. Time-budgets of Camargue horses. II. Time-
 amount of digesta in the reticulo-rumen and the rate of
                                 budgets of adult horses and weaned subadults. Behaviour
 disappearance of digesta from the alimentary tract. British
                                 72:26–49.
 Journal of Nutrition 15:531–540.               Edwards, D. C., and J. D. Huebner. 1977. Feeding and growth
Caraco, T. 1979. Time budgeting and group size: a test of     rates of Polinices duplicatus preying on Mya arenaria at
 theory. Ecology 60:618–627.                   Barnstable Harbor, Massachusetts. Ecology 58:1218–1236.
Casas, J., W. S. C. Gurney, R. Nisbet, and O. Roux. 1993.    Elliott, J. M. 1983. The responses of the aquatic parasitoid
 A probabilistic model for the functional response of a par-   Agriotypus armatus (Hymenoptera: Agriotypidae) to the
 asitoid at the behavioural time-scale. Journal of Animal    spatial distribution and density of its caddis host Silo pal-
 Ecology 62:194–204.                       lipes (Trichoptera: Goeridae). Journal of Animal Ecology
Christoffersen, K., and A.-M. Jespersen. 1986. Gut evacu-     52:315–330.
 ation rates and ingestion rates of Eudiaptomus graciloides  Elliott, J. M., and L. Persson. 1978. The estimation of daily
 measured by means of gut fluorescence method. Journal of     rates of food consumption for fish. Journal of Animal Ecol-
 Plankton Research 8:973–983.                  ogy 47:977–991.
Cook, R. M., and B. J. Cockrell. 1978. Predator ingestion    Engen, S., and N. C. Stenseth. 1984. A general version of
 rate and its bearing on feeding time and the theory of op-   optimal foraging theory: the effect of simultaneous en-
 timal diets. Journal of Animal Ecology 47:529–547.       counters. Theoretical Population Biology 26:192–204.
Cosner, C., D. L. DeAngelis, J. S. Ault, and D. B. Olson.    Farnsworth, K. D., and A. W. Illius. 1996. Large grazers back
 1999. Effects of spatial grouping on the functional re-     in the fold: generalizing the prey model to incorporate
 sponse of predators. Theoretical Population Biology 56:     mammalian herbivores. Functional Ecology 10:678–680.
 65–75.                            Farnsworth, K. D., and A. W. Illius. 1998. Optimal diet choice
Crisp, D. J., Yule, A. B., and K. N. White. 1985. Feeding     for large herbivores: an extended contingency model. Func-
 by oyster larvae: the functional response, energy budget    tional Ecology 12:74–81.
 and a comparison with mussel larvae. Journal of the Marine  Fedorenko, A. 1975. Feeding characteristics and predation
 Biological Association of the United Kingdom 65:759–      impact of Chaoborus (Diptera) larvae in a small lake. Lim-
 784.                              nology and Oceanography 20:250–258.
Crowley, P. H. 1973. Filtering rate inhibition of Daphnia    Fenchel, T. 1982a. Ecology of heterotrophic microflagellates.
 pulex in Wintergreen Lake water. Limnology and Ocean-      I. Some important forms and their functional morphology.
 ography 18:394–402.                       Marine Ecology Progress Series 8:211–223.
                    PREDATOR FUNCTIONAL RESPONSES
Febraury 2002                                                        109

Fenchel, T. 1982b. Ecology of heterotrophic microflagellates.    constrained processes. Journal of Theoretical Biology 176:
 II. Bioenergetics and growth. Marine Ecology Progress Se-     33–37.
 ries 8:225–231.                        Herbers, J. 1981. Time resources and laziness in animals.
Fleschner, C. A. 1950. Studies on searching capacity of the     Oecologia 49:252–262.
 larvae of three predators of the citrus red mite. Hilgardia  Hirakawa, H. 1995. Diet optimization with a nutrient or toxin
 20:233–265.                            constraint. Theoretical Population Biology 47:331–346.
Forchhammer, M. C., and J. J. Boomsma. 1995. Foraging      Hirakawa, H. 1997a. How important is digestive quality? A
 strategies and seasonal diet optimization of muskoxen in     correction of Verlinden and Wiley’s digestive rate model.
 West Greenland. Oecologia 104:169–180.              Evolutionary Ecology 11:249–252.
Formanowicz, D. R., Jr. 1984. Foraging tactics of an aquatic   Hirakawa, H. 1997b. Digestion-constrained optimal foraging
 insect: partial consumption of prey. Animal Behaviour 32:     in generalist mammalian herbivores. Oikos 78:37–47.
 774–781.                            Hixon, M. A., F. L. Carpenter, and D. C. Paton. 1983. Ter-
Frost, B. W. 1972. Effects of size and concentration of food    ritory area, flower density, and time budgeting in hum-
 particles on the feeding behavior of the marine planktonic    mingbirds: an experimental and theoretical analysis. Amer-
 copepod Calanus pacificus. Limnology and Oceanography       ican Naturalist 122:366–391.
 17:805–815.                          Holling, C. S. 1959a. The components of predation as re-
Fryxell, J. M. 1991. Forage quality and aggregation by large    vealed by a study of small-mammal predation of the Eu-
 herbivores. American Naturalist 138:478–498.           ropean pine sawfly. Canadian Entomologist 91:293–320.
Fryxell, J. M., and P. Lundberg. 1997. Individual behavior    Holling, C. S. 1959b. Some characteristics of simple types
 and community dynamics. Chapman and Hall, New York,        of predation and parasitism. Canadian Entomologist 91:
 New York, USA.                          385–398.
Fryxell, J. M., S. M. Vamosi, R. A. Walton, and C. M. Doucet.  Holling, C. S. 1965. The functional response of predators to
 1994. Retention time and the functional response of bea-     prey density. Memoirs of the Entomological Society of
 vers. Oikos 71:207–214.                      Canada 45:1–60.
Fujii, K., C. S. Holling, and P. M. Mace. 1986. A simple     Holling, C. S. 1966. The functional response of invertebrate
 generalized model of attack by predators and parasites.      predators to prey density. Memoirs of the Entomological
 Ecological Research 1:141–156.                  Society of Canada 48:1–86.
Gause, G. F. 1934. The struggle for existence. Williams and   Holling, C. S., and S. Buckingham. 1976. A behavioral model
 Wilkins, Baltimore, Maryland, USA.                of predator prey functional responses. Behavioral Science
Geller, W. 1975. Die Nahrungsaufnahme von Daphnia pulex       21:183–195.
 in Abhangigkeit von der Futterkonzentration, der Temper-
     ¨                           Holmgren, N. 1995. The ideal free distribution of unequal
 atur, der Korpergroße und dem Hungerzustand der Tiere.
        ¨   ¨                       competitors: predictions from a behaviour-based functional
 Archiv fur Hydrobiologie Supplement 48:47–107.
      ¨                            response. Journal of Animal Ecology 64:197–212.
Gerritsen, J., and J. R. Strickler. 1977. Encounter probabil-  Illius, A. W., and I. J. Gordon. 1991. Prediction of intake
 ities and community structure in zooplankton: a mathe-      and digestion in ruminants by a model of rumen kinetics
 matical model. Journal of the Fisheries Research Board of     integrating animal size and plant characteristics. Journal of
 Canada 34:73–82.                         Agricultural Science 116:145–157.
Getty, T., and H. R. Pulliam. 1991. Random prey detection    Inman, A. J., and J. Krebs. 1987. Predation and group living.
 with pause-travel search. American Naturalist 138:1459–      Trends in Ecology and Evolution 2:31–32.
 1477.                             Ivlev, V. S. 1961. Experimental ecology of the feeding of
Giguere, L. A., A. Delage, L. M. Dill, and J. Gerritsen. 1982.
   `         ˆ                      fishes. Yale University Press, New Haven, USA.
 Predicting encounter rates for zooplankton: a model as-    Jeschke, J. M., and R. Tollrian. 2000. Density-dependent ef-
 suming a cylindrical encounter field. Canadian Journal of     fects of prey defences. Oecologia 123:391–396.
 Fisheries and Aquatic Sciences 39:237–242.           Johnson, D. M., B. G. Akre, and P. H. Crowley. 1975. Mod-
Griffiths, K. J., and C. S. Holling. 1969. A competition sub-    eling arthropod predation: wasteful killing by damselfly
 model for parasites and predators. Canadian Entomologist     naiads. Ecology 56:1081–1093.
 101:785–818.                          Jonsson, P. R., and P. Tiselius. 1990. Feeding behaviour, prey
                                  detection and capture efficiency of the copepod Acartia
Halbach, U., and G. Halbach-Keup. 1974. Quantitative Be-
                                  tonsa feeding on planktonic ciliates. Marine Ecology Pro-
 ziehungen zwischen Phytoplankton und der Populations-
 dynamik des Rotators Brachionus calyciflorus Pallas. Be-      gress Series 60:35–44.
 funde aus Laboratoriumsexperimenten und Freilandunter-     Juliano, S. A. 1989. Queueing models of predation and the
 suchungen. Archiv fur Hydrobiologie 73:273–309.
             ¨                     importance of contingent behavioural choices for optimal
Hansen, B., K. S. Tande, and U. C. Berggreen. 1990. On the     foragers. Animal Behaviour 38:757–770.
 trophic fate of Phaeocystis pouchetii (Hariot). III. Func-   Kaiser, J. 1998. Sea otter declines blamed on hungry killers.
 tional responses in grazing demonstrated on juvenile stages    Science 282:390–391.
 of Calanus finmarchicus (Copepoda) fed diatoms and       Karban, R., and I. T. Baldwin. 1997. Induced responses to
 Phaeocystis. Journal of Plankton Research 12:1173–1187.      herbivory. University of Chicago Press, Chicago, USA.
Hassell, M. P., J. H. Lawton, and J. R. Beddington. 1976.    Kenward, R. E., and R. M. Sibly. 1977. A woodpigeon (Co-
                                  lumba palumbus) feeding preference explained by a diges-
 The components of arthropod predation. 1. The prey death
 rate. Journal of Animal Ecology 45:135–164.            tive bottle-neck. Journal of Applied Ecology 14:815–826.
Hassell, M. P., J. H. Lawton, and J. R. Beddington. 1977.    Kersten, M., and W. Visser. 1996. The rate of food processing
 Sigmoid functional responses by invertebrate predators and    in the oystercatcher: food intake and energy expenditure
 parasitoids. Journal of Animal Ecology 46:249–262.        constrained by a digestive bottleneck. Functional Ecology
Hassell, M. P., and G. C. Varley. 1969. New inductive pop-     10:440–448.
 ulation model for insect parasites and its bearing on bio-   Kitchell, J. A., C. H. Boggs, J. F. Kitchell, and J. A. Rice.
 logical control. Nature 223:1133–1136.              1981. Prey selection by naticid gastropods: experimental
Head, E. J. H. 1986. Estimation of Arctic copepod grazing      tests and application to the fossil record. Paleobiology 7:
 rates in vivo and comparison with in-vitro methods. Marine    533–552.
 Biology 92:371–379.                      Koopman, B. O. 1956. The theory of search. I. Kinematic
Henson, S. M., and T. G. Hallam. 1995. Optimal feeding via     bases. Operations Research 4:324–346.
                                                      Ecological Monographs
                       JONATHAN M. JESCHKE ET AL.
110
                                                          Vol. 72, No. 1

Korona, R. 1989. Ideal free distribution of unequal compet-    Milinski, V. M., and E. Curio. 1975. Untersuchungen zur
 itors can be determined by the form of competition. Journal    Selektion durch Rauber gegen Vereinzelung der Beute.
                                             ¨
 of Theoretical Biology 138:347–352.                Zeitschrift fur Tierpsychologie 37:400–402.
                                         ¨
Laca, E. A., E. D. Ungar, and M. W. Demment. 1994. Mech-     Miller, R. C. 1922. The significance of the gregarious habit.
 anisms of handling time and intake rate of a large mam-      Ecology 3:375–382.
 malian grazer. Applied Animal Behaviour Science 39:3–      Mills, N. J. 1982. Satiation and the functional response: a
 19.                                test of a new model. Ecological Entomology 7:305–315.
Lam, R. K., and B. W. Frost. 1976. Model of copepod filtering   Mitchell, W. A., and J. S. Brown. 1990. Density dependent
 response to changes in size and concentration of food. Lim-    harvest rates by optimal foragers. Oikos 57:180–190.
 nology and Oceanography 21:490–500.               Moncorps, S., P. Bousses, D. Reale, and J.-L. Chapuis. 1997.
                                              `    ´
                                  Diurnal time budget of the mouflon (Ovis musimon) on the
Landeau, L., and J. Terborgh. 1986. Oddity and the ‘‘con-
 fusion effect’’ in predation. Animal Behaviour 34:1372–      Kerguelen archipelago: influence of food resources, age,
 1380.                               and sex. Canadian Journal of Zoology 75:1828–1834.
Lazarus, J. 1979. The early warning function of flocking in    Morgan, M. J., and J.-G. J. Godin. 1985. Antipredator ben-
 birds: an experimental study with captive Quelea. Animal     efits of schooling behaviour in a cyprinodontid fish, the
                                  barred killifish (Fundulus diaphanus). Zeitschrift fur Tierp-
 Behaviour 27:855–865.                                                 ¨
Lehman, J. T. 1976. The filter-feeder as an optimal forager,     sychologie 70:236–246.
 and the predicted shapes of feeding curves. Limnology and    Mori, H., and D. A. Chant. 1966. The influence of prey den-
 Oceanography 21:501–516.                     sity, relative humidity, and starvation on the predacious
                                  behavior of Phytoseiulus persimilis Athias-Henriot (Aca-
Leonardsson, K., and F. Johansson. 1997. Optimum search
 speed and activity: a dynamic game in a three-link trophic    rina: Phytoseiidae). Canadian Journal of Zoology 44:483–
 system. Journal of Evolutionary Biology 10:703–729.        491.
Longstaff, B. C. 1980. The functional response of a predatory   Murtaugh, P. A. 1984. Variable gut residence time: problems
 mite and the nature of the attack rate. Australian Journal    in inferring feeding rate from stomach fullness of a mysid
 of Ecology 5:151–158.                       crustacean. Canadian Journal of Fisheries and Aquatic Sci-
Lucas, J. R. 1985. Partial prey consumption by antlion larvae.   ences 41:1287–1293.
 Animal Behaviour 33:945–958.                  Nakamura, K. 1974. A model of the functional response of
Lucas, J. R., and A. Grafen. 1985. Partial prey consumption     a predator to prey density involving the hunger effect. Oec-
 by ambush predators. Journal of Theoretical Biology 113:     ologia 16:265–278.
 455–473.                            Neill, S. R. S. J., and J. M. Cullen. 1974. Experiments on
Lundberg, P., and K. Danell. 1990. Functional responses of     whether schooling by their prey affects the hunting behav-
 browsers: tree exploitation by moose. Oikos 58:378–384.      iour of cephalopods and fish predators. Journal of Zoology
Lundberg, P., and M. Ostrom. 1990. Functional responses of
               ¨                   172:549–569.
 optimally foraging herbivores. Journal of Theoretical Bi-    Nelmes, A. J. 1974. Evaluation of the feeding behaviour of
                                  Prionchulus punctatus (Cobb), a nematode predator. Jour-
 ology 144:367–377.
Maxson, S. J., and L. W. Oring. 1980. Breeding season time     nal of Animal Ecology 43:553–565.
 and energy budgets of the polyandrous spotted sandpiper.    Norris, K., and I. Johnstone. 1998. Interference competition
 Behaviour 74:200–263.                       on the functional response of oystercatchers searching for
May, R. M. 1978. Host–parasitoid systems in patchy envi-      cockles by touch. Animal Behaviour 56:639–650.
 ronments: a phenomenological model. Journal of Animal      Oaten, A. 1977. Transit time and density-dependent predation
 Ecology 47:833–844.                        on a patchily-distributed prey. American Naturalist 111:
Mayzaud, P., and S. A. Poulet. 1978. The importance of the     1061–1075.
 time factor in the response of zooplankton to varying con-   Oaten, A., and W. W. Murdoch. 1975. Switching, functional
 centrations of naturally occurring particulate matter. Lim-    response, and stability in predator–prey systems. American
 nology and Oceanography 23:1144–1154.               Naturalist 109:299–318.
McMahon, J. W., and F. H. Rigler. 1963. Mechanisms reg-      Owen-Smith, N. 1994. Foraging responses of kudus to sea-
 ulating the feeding rate of Daphnia magna Straus. Canadian    sonal changes in food resources: elasticity in constraints.
 Journal of Zoology 41:321–332.                  Ecology 75:1050–1062.
McNair, J. N. 1980. A stochastic foraging model with pred-    Owen-Smith, N., and P. Novellie. 1982. What should a clever
 ator training effects: I. Functional response, switching, and   ungulate eat? American Naturalist 119:151–178.
 run lengths. Theoretical Population Biology 17:141–166.     Packer, C., and L. Ruttan. 1988. The evolution of cooperative
McNair, J. N. 1983. A class of patch-use strategies. American    hunting. American Naturalist 132:159–198.
 Zoologist 23:303–313.                      Paffenhofer, G.-A., J. R. Strickler, and M. Alcaraz. 1982.
                                     ¨
McNamara, J. M., and A. I. Houston. 1994. The effect of a      Suspension-feeding by herbivorous calanoid copepods: a
 change in foraging options on intake rate and predation      cinematographic study. Marine Biology 67:193–199.
 rate. American Naturalist 144:978–1000.             Paloheimo, J. E. 1971a. On a theory of search. Biometrika
Metz, J. A. J., M. W. Sabelis, and J. H. Kuchlein. 1988.      58:61–75.
 Sources of variation in predation rates at high prey den-    Paloheimo, J. E. 1971b. A stochastic theory of search: im-
 sities: an analytic model and a mite example. Experimental    plications for predator–prey situations. Mathematical Bio-
 and Applied Acarology 5:187–205.                 sciences 12:105–132.
Metz, J. A. J., and F. H. D. van Batenburg. 1985a. Holling’s   Parker, G. A., and W. J. Sutherland. 1986. Ideal free distri-
 ‘‘hungry mantid’’ model for the invertebrate functional re-    butions when individuals differ in competitive ability: phe-
 sponse considered as a Markov process. Part I: The full      notype limited ideal free models. Animal Behaviour 34:
 model and some of its limits. Journal of Mathematical Bi-     1222–1242.
 ology 22:209–238.                        Parsons, A. J., J. H. M. Thornley, J. Newman, and P. D.
Metz, J. A. J., and F. H. D. van Batenburg. 1985b. Holling’s    Penning. 1994. A mechanistic model of some physical
 ‘‘hungry mantid’’ model for the invertebrate functional re-    determinants of intake rate and diet selection in a two-
 sponse considered as a Markov process. Part II: Negligible    species temperate grassland sward. Functional Ecology 8:
 handling time. Journal of Mathematical Biology 22:239–      187–204.
 257.                              Parsons, T. R., R. J. LeBrasseur, and J. D. Fulton. 1967. Some
                    PREDATOR FUNCTIONAL RESPONSES
Febraury 2002                                                       111

 observations on the dependence of zooplankton grazing on    foraging in mammalian herbivores: new models of func-
 the cell size and concentration of phytoplankton blooms.    tional response. American Naturalist 140:325–348.
 Journal of the Oceanographical Society of Japan 23:10–    Stephens, D. W., and J. R. Krebs. 1986. Foraging theory.
 17.                               Princeton University Press, Princeton, New Jersey, USA.
Parsons, T. R., R. J. LeBrasseur, J. D. Fulton, and O. D.    Streams, F. A. 1994. Effect of prey size on attack components
 Kennedy. 1969. Production studies in the strait of Georgia.   of the functional response by Notonecta undulata. Oecol-
 Part II. Secondary production under the Fraser River plume,   ogia 98:57–63.
 February to May, 1967. Journal of Experimental Marine     Thompson, D. J. 1978. Towards a realistic predator–prey
 Biology and Ecology 3:39–50.                  model: the effect of temperature on the functional response
Rao, C. R., and A. M. Kshirsagar. 1978. A semi-Markovian     and life history of larvae of the damselfly Ischnura elegans.
 model for predator–prey interactions. Biometrics 34:611–    Journal of Animal Ecology 47:757–767.
 619.                             Tollrian, R., and C. D. Harvell, editors. 1999. The ecology
Rashevsky, N. 1959. Some remarks on the mathematical the-     and evolution of inducible defenses. Princeton University
 ory of nutrition of fishes. Bulletin of Mathematical Bio-    Press, Princeton, New Jersey, USA.
 physics 21:161–182.                      Tostowaryk, W. 1972. The effects of prey defense on the
Real, L. A. 1977. The kinetics of functional response. Amer-   functional response of Podisus modestus to densities of the
 ican Naturalist 111:289–300.                  sawflies Neodiprion swainei and N. pratti banksianae (Hy-
Real, L. A. 1979. Ecological determinants of functional re-
                                 menoptera: Neodiprionidae). Canadian Entomologist 104:
 sponse. Ecology 60:481–485.
                                 61–69.
Rigler, F. H. 1961. The relationship between concentration
                                Triplet, P., R. A. Stillman, and J. D. Goss-Custard. 1999. Prey
 of food and feeding rate of Daphnia magna Straus. Ca-
                                 abundance and the strength of interference in a foraging
 nadian Journal of Zoology 39:857–868.
                                 shorebird. Journal of Animal Ecology 68:254–265.
Rogers, D. 1972. Random search and insect population mod-
                                Trudell, J., and R. G. White. 1981. The effect of forage struc-
 els. Journal of Animal Ecology 41:369–383.
                                 ture and availability on food intake, biting rate, bite size
Rogowitz, G. L. 1997. Locomotor and foraging activity of
                                 and daily eating time of reindeer. Journal of Applied Ecol-
 the white-tailed jackrabbit (Lepus townsendii). Journal of
                                 ogy 18:63–82.
 Mammalogy 78:1172–1181.
                                Ungar, E. D., and I. Noy-Meir. 1988. Herbage intake in re-
Rothley, K. D., O. J. Schmitz, and J. L. Cohon. 1997. For-
                                 lation to availability and sward structure: grazing processes
 aging to balance conflicting demands: Novel insights from
                                 and optimal foraging. Journal of Applied Ecology 25:1045–
 grasshoppers under predation risk. Behavioral Ecology 8:
                                 1062.
 551–559.
                                Verlinden, C., and R. H. Wiley. 1989. The constraints of
Royama, T. 1971. A comparative study of models for pre-
                                 digestive rate: an alternative model of diet selection. Evo-
 dation and parasitism. Researches on Population Ecology
                                 lutionary Ecology 3:264–273.
 S1:1–90.
Ruxton, G. D., and W. S. C. Gurney. 1994. Deriving the     Vermeij, G. J. 1982. Unsuccessful predation and evolution.
 functional response without assuming homogeneity. Amer-     American Naturalist 120:701–720.
 ican Naturalist 144:537–541.                 Visser, M., and L. J. Reinders. 1981. Waiting time as a new
Ruxton, G. D., W. S. C. Gurney, and A. M. deRoos. 1992.      component in functional response models. Netherlands
 Interference and generation cycles. Theoretical Population   Journal of Zoology 31:315–328.
 Biology 42:235–253.                      Wanink, J., and L. Zwarts. 1985. Does an optimally foraging
Saariko, J., and I. Hanksi. 1990. Timing of rest and sleep in   oystercatcher obey the functional response? Oecologia 67:
 foraging shrews. Animal Behaviour 40:861–869.          98–106.
Schmitz, O. J. 1995. Functional responses of optimal con-    Ward, D. 1992. The role of satisficing in foraging theory.
 sumers and the implication for regulation of resource pop-   Oikos 63:312–317.
 ulations. Wildlife Research 22:101–111.            Ward, D. 1993. Foraging theory, like all other fields of sci-
Schmitz, O. J., A. P. Beckerman, and S. Litman. 1997. Func-    ence, needs multiple working hypotheses. Oikos 67:376–
 tional responses of adaptive consumers and community sta-    378.
 bility with emphasis on the dynamics of plant–herbivore    Watt, K. E. F. 1959. A mathematical model for the effect of
 systems. Evolutionary Ecology 11:773–784.
                                 densities of attacked and attacking species on the number
Schoener, T. W. 1971. Theory of feeding strategies. Annual
                                 attacked. Canadian Entomologist 91:129–144.
 Review of Ecology and Systematics 2:370–404.
                                Weiner, J. 1992. Physiological limits to sustainable energy
Shipley, L. A., J. E. Gross, D. E. Spalinger, N. T. Hobbs,
                                 budgets in birds and mammals: ecological implications.
 and B. A. Wunder. 1994. The scaling of intake rate in
                                 Trends in Ecology and Evolution 7:384–388.
 mammalian herbivores. American Naturalist 143:1055–
                                Werner, E. E., and B. R. Anholt. 1993. Ecological conse-
 1082.
                                 quences of the tradeoff between growth and mortality
Sih, A. 1980. Optimal foraging: partial consumption of prey.
                                 rates mediated by plasticity. American Naturalist 142:
 American Naturalist 116:281–290.
                                 242–272.
Sih, A. 1984. Optimal behavior and density-dependent pre-
                                Williamson, C. E. 1984. Laboratory and field experiments
 dation. American Naturalist 123:314–326.
                                 on the feeding ecology of the cyclopoid copepod, Meso-
Sih, A., G. Englund, and D. Wooster. 1998. Emergent impacts
                                 cyclops edax. Freshwater Biology 14:575–585.
 of multiple predators on prey. Trends in Ecology and Evo-
                                Wilmshurst, J. F., J. M. Fryxell, and C. M. Bergman. 2000.
 lution 13:350–355.
                                 The allometry of patch selection in ruminants. Proceedings
Sjoberg, S. 1980. Zooplankton feeding and queueing theory.
 ¨
                                 of The Royal Society of London Series B 267:345–349.
 Ecological Modelling 10:215–225.
                                Wilmshurst, J. F., J. M. Fryxell, and P. E. Colucci. 1999. What
Solomon, M. E. 1949. The natural control of animal popu-
                                 constrains daily intake in Thomson’s gazelles? Ecology 80:
 lations. Journal of Animal Ecology 18:1–35.
                                 2338–2347.
Soluk, D. A. 1993. Multiple predator effects: predicting com-
                                Wilmshurst, J. F., J. M. Fryxell, and R. J. Hudson. 1995.
 bined functional response of streamfish and invertebrate
                                 Forage quality and patch choice by wapiti (Cervus elaphus).
 predators. Ecology 74:219–225.
Spalinger, D. E., and N. T. Hobbs. 1992. Mechanisms of      Behavioral Ecology 6:209–217.
                                                                       Ecological Monographs
                          JONATHAN M. JESCHKE ET AL.
112
                                                                           Vol. 72, No. 1

                      APPENDIX A: FROM EQ. 12 TO EQ. 13
  Solving Eq. 12 for y(x) gives the two solutions y1(x) and y2(x): lim y2 (x) 0,                                       (A1)
                                                  x→0

                                                 As the limits indicate, only the second solution, y2(x),
                                       c) 2 ]
         ax(b  c)       ax[2(b   c)   ax(b
     1            1
y1 (x)                                        ,    makes sense biologically. However, y2(x) is not defined for a
                   2abcx                            0, b   0, c  0, or x   0. Eq. 12 helps to find the cor-
  lim y1 (x) →   ,                                     responding equations: For a    0 or x  0, Eq. 12 gives y(x)
  x→0
                                                 0; for b   0, Eq. 12 gives y(x)        acx); for c
                                                                    (ac)/(1
                                       c) 2 ]
         ax(b  c)       ax[2(b   c)   ax(b
     1            1                               0, Eq. 12 gives y(x)         abx); and for b   c
                                                             (ab)/(1               0,
y2 (x)                                        ,
                                                Eq. 12 gives y(x)   ax.
                   2abcx


                          APPENDIX B
 An estimation of Polinices duplicatus (Gastropoda: Naticacea) handling time feeding on Mya arenaria (Bivalvia) in the
year-round experiments of Edwards and Huebner (1977).

               Mya lengthi      Mya shell                 Polinices drilling                 No. preyed
  Mya size class i                                                         bi[d]¶    ni (
               [ 5 mm]†      thicknessi [mm]‡                  timei [d]§                      385)†
                                                               ˆi  ˆ
       1          15           0.121                      0.226        0.95  0.458       73
       2          25           0.277                      0.517        0.95  1.048       160
       3          35           0.433                      0.809        0.95  1.639       108
       4          45           0.589                      1.100        0.95  2.229       26
       5          55           0.745                      1.392        0.25  4.871       16
       6          65           0.901                      1.683        0.05  19.355        2
                                   6
                              1
  Resulting mean estimated handling time b                  bi ni      1.4 d.
                     ˆ                  ˆ
                              385  i    1
 † Table 1 in Edwards and Huebner (1977).
 ‡ Table 3 in Kitchell et al. (1981): M. arenaria shell thicknessi (mm)     0.113  0.0156   lengthi (mm).
 § Kitchell et al. (1981): P. duplicatus drilling time   1.868 d/mm.
  According to Kitchell et al. (1981), attack efficiency of P. duplicatus mainly depends on predator and prey size. Given
a certain predator size, is almost unity for prey below a critical size and almost zero for prey beyond that critical size.
The critical size for M. arenaria is given in Fig. 7 in Kitchell et al. (1981). The predator sizes are given in Table 3 in Edwards
and Huebner (1977). Edwards and Huebner used four individual predators with mean sizes in the relevant period (14 June–
29 August, where maximum predation rate ymax     0.63 M. arenaria/d have been reported for P. duplicatus) of 37.9 mm, 41
mm, 42.15 mm, and 50.45 mm, respectively. The corresponding critical M. arenaria lengths are roughly 53 mm for the three
small predator individuals and 60 mm for the largest one. Therefore, M. arenaria of size classes 1, 2, 3, and 4 could be
                                      0.95), M. arenaria of size class 5 could basically only
easily attacked by all four predator individuals (ˆ 1 ˆ 2 ˆ 3    ˆ4
                          0.25), and M. arenaria of size class 6 could only hardly be attacked by all
be attacked by one of the four predators ( ˆ 5
four predators ( ˆ 6  0.05).
 ¶ Estimated P. duplicatus handling time (for M. arenaria size class i):
                                                ) 1]
                            bi [d]    [1.5         (2       drilling timei .                  (B1)
                            ˆ                    i

Derivation: From Eq. 4, bi [d]   (drilling time i / i) eating time i  (drilling time i i)  drilling time i (Kitchell et al.
              ˆ
1981). However, this calculation overestimates handling time, because it is based on the assumption that unsuccessful
drills last as long as successful ones. Assuming that unsuccessful drills last, on average, half the time of successful ones,
leads to Eq. B1.
by Mark Novak last modified 17-10-2006 11:12
 

Built with Plone