Skip to main content
Log in

Ecological constraints on the origin of neurones

  • Published:
Journal of Mathematical Biology Aims and scope Submit manuscript

Abstract

The basic functional characteristics of spiking neurones are remarkably similar throughout the animal kingdom. Their core design and function features were presumably established very early in their evolutionary history. Identifying the selection pressures that drove animals to evolve spiking neurones could help us interpret their design and function today. This paper provides a quantitative argument, based on ecology, that animals evolved neurones after they started eating each other, about 550 million years ago. We consider neurones as devices that aid an animal’s foraging performance, but incur an energetic cost. We introduce an idealised stochastic model ecosystem of animals and their food, and obtain an analytic expression for the probability that an animal with a neurone will fix in a neurone-less population. Analysis of the fixation probability reveals two key results. First, a neurone will never fix if an animal forages low-value food at high density, even if that neurone incurs no cost. Second, a neurone will fix with high probability if an animal is foraging high-value food at low density, even if that neurone is expensive. These observations indicate that the transition from neurone-less to neurone-armed animals can be facilitated by a transition from filter-feeding or substrate grazing to episodic feeding strategies such as animal-on-animal predation (macrophagy).

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

Access this article

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

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4

Similar content being viewed by others

Notes

  1. These general assumptions also guarantee absorption in finite time.

  2. This can be shown formally by using the procedure outlined by Wald (1944) to obtain the conditional characteristic functions of the time to absorption, then calculating conditional expected times from the characteristic functions.

  3. If we disallow food items to be located inside the decider, then we get rid of infinitely strong signals at \(r=0\) and the integrals converge.

References

  • Arendt D, Denes AS, Jekely G, Tessmar-Raible K (2008) The evolution of nervous system centralization. Philos Trans R Soc B 363:1523–1528

    Article  Google Scholar 

  • Aris-Brosou S, Yang Z (2003) Bayesian models of episodic evolution support a late Precambrian explosive diversification of the metazoa. Mol Biol Evol 20(12):1947–1954

    Article  Google Scholar 

  • Ball EE, Hayward DC, Saint R, Miller DJ (2004) A simple plan—cnidarians and the origins of developmental mechanisms. Nat Rev Gen 5(8):567–577

    Article  Google Scholar 

  • Barnett N (1975) Application of a generalisation of Wald’s identity to solving certain multi-stream storage problems. J Hydrol 26:95–113

    Article  Google Scholar 

  • Bengtson S, Zhao Y (1992) Predatorial borings in late Precambrian mineralized exoskeletons. Science 257:367–369

    Article  Google Scholar 

  • Berg HC (1975) Chemotaxis in bacteria. Annu Rev Biophys Bio 4(1):119–136

    Article  Google Scholar 

  • Bottjer DJ, Clapham ME (2006) Evolutionary paleoecology of Ediacaran benthic marine animals, chap. 4, Springer, pp 91–114

  • Broom M, Ruxton GD (2005) You can run- or you can hide: optimal strategies for cryptic prey against pursuit predators. Behav Ecol 16(3):534–540

    Article  Google Scholar 

  • Budd GE, Jensen S (2000) A critical reappraisal of the fossil record of the bilaterian phyla. Biol Rev 75(2):253–295

    Article  Google Scholar 

  • Chen JY, Oliveri P, Gao F, Dornbos SQ, Li CW, Bottjer DJ, Davidson EH (2002) Precambrian animal life: probable developmental and adult cnidarian forms from southwest China. Dev Biol 248(1):182–196

    Article  Google Scholar 

  • Chen Z, Zhou C, Meyer M, Xiang K, Schiffbauer JD, Yuan X, Xiao S (2013) Trace fossil evidence for Ediacaran bilaterian animals with complex behaviors. Precambr Res 224:690–701

    Article  Google Scholar 

  • Chittka L, Skorupski P, Raine NE (2009) Speed–accuracy tradeoffs in animal decision making. Trends Ecol Evol 24(7):400–407

    Article  Google Scholar 

  • Clapham ME, Narbonne GM, Gehling JG (2003) Paleoecology of the oldest known animal communitites: Ediacaran assemblages at Mistaken Point, Newfoundland. Paleobiology 29(4):527–544

    Article  Google Scholar 

  • Clarkson E, Levi-Setti R, Horvath G (2006) The eyes of trilobites: the oldest preserved visual system. Arthropod Struct Dev 35(4):247–259

    Article  Google Scholar 

  • Crotty P, Sangrey T, Levy WB (2006) Metabolic energy cost of action potential velocity. J Neurophysiol 96(3):1237–1246

    Article  Google Scholar 

  • Dawkins R, Krebs JR (1979) Arms races between and within species. Proc R Soc B 205:489–511

    Article  Google Scholar 

  • Denes AS, Jekely G, Steinmetz PR, Raible F, Snyman H, Prud’homme B, Ferrier DE, Balavoine G, Arendt D (2007) Molecular architecture of annelid nerve cord supports common origin of nervous system centralization in bilateria. Cell 129(2):277–288

    Article  Google Scholar 

  • Droser ML, Gehling JG, Jensen SR (2006) Assemblage palaeoecology of the Ediacara biota: the unabridged edition? Palaeogeogr Palaeoclimatol 232(2):131–147

    Article  Google Scholar 

  • Elliott GRD, Leys SP (2007) Coordinated contractions effectively expel water from the aquiferous system of a freshwater sponge. J Exp Biol 210(21):3736–3748

    Article  Google Scholar 

  • Erwin DH, Laflamme M, Tweedt SM, Sperling EA, Pisani D, Peterson KJ (2011) The Cambrian conundrum: early divergence and later ecological success in the early history of animals. Science 334:1091–1097

  • Ettinger-Epstein P, Whalan S, Battershill CN, de Nys R (2008) A hierarchy of settlement cues influences larval behaviour in a coral reef sponge. Mar Ecol Prog Ser 365:103–113

    Article  Google Scholar 

  • Ewens WJ (2004) Mathematical population genetics I. Theoretical introduction. Springer, New York

    Book  MATH  Google Scholar 

  • Fedonkin MA (1997) The Late Precambrian fossil Kimberella is a mollusc-like bilaterian organism. Nature 388:868–871

    Article  Google Scholar 

  • Galliot B, Quiquand M (2011) A two-step process in the emergence of neurogenesis. Eur J Neurosci 34(6):847–862

    Article  Google Scholar 

  • Galliot B, Quiquand M, Ghila L, de Rosa R, Miljkovic-Licina M, Chera S (2009) Origins of neurogenesis, a cnidarian view. Dev Biol 332:2–24

    Article  Google Scholar 

  • Ghysen A (2003) The origin and evolution of the nervous system. Int J Dev Biol 47(7–8):555–562

    Google Scholar 

  • Gomez-Skarmeta JL, Campuzano S, Modolell J (2003) Half a century of neural prepatterning: the story of a few bristles and many genes. Nat Rev Neurosci 4(7):587–598

    Article  Google Scholar 

  • Hayakawa E, Fujisawa C, Fujisawa T (2004) Involvement of Hydra achaete–scute gene CnASH in the differentiation pathway of sensory neurons in the tentacles. Dev Genes Evol 214(10):486–492

    Google Scholar 

  • Hirth F (2010) On the origin and evolution of the tripartite brain. Brain Behav Evol 76(1):3–10

    Article  Google Scholar 

  • Holland ND (2003) Early central nervous system evolution: an era of skin brains? Nat Rev Neurosci 4(8):617–627

    Article  Google Scholar 

  • Hua H, Pratt BR, Zhang LY (2003) Borings in Cloudina shells: complex predator-prey dynamics in the terminal Neoproterozoic. PALAIOS 18(4/5):454–459

    Article  Google Scholar 

  • Jacobs D, Nakanishi N, Yuan D, Camara A, Nichols SA, Hartenstein V (2007) Evolution of sensory structures in basal metazoa. Integr Comp Biol 47(5):712–723

    Article  Google Scholar 

  • Jensen S (2003) The Proterozoic and earliest Cambrian trace fossil record; patterns, problems and perspectives. Integr Comp Biol 43(1):219–228

    Article  Google Scholar 

  • Jensen S, Droser ML, Gehling JG (2005) Trace fossil preservation and the early evolution of animals. Palaeogeogr Palaeoclimatol 220:19–29

    Article  Google Scholar 

  • Jensen S, Saylor BZ, Gehling JG, Germs GJB (2000) Complex trace fossils from the terminal Proterozoic of Namibia. Geology 28(2):143–146

    Article  Google Scholar 

  • Laflamme M, Narbonne GM (2008) Ediacaran fronds. Palaeogeogr Palaeoclimatol 258(3):162–179

    Article  Google Scholar 

  • Laughlin SB (2001) Energy as a constraint on the coding and processing of sensory information. Curr Opin Neurobiol 11(4):475–480

    Article  Google Scholar 

  • Laughlin SB, de Ruyter van Steveninck RR, Anderson JC (1998) The metabolic cost of neural information. Nat Neurosci 1(1):36–41

    Article  Google Scholar 

  • Leys SP, Cronin TW, Degnan BM, Marshall JN (2002) Spectral sensitivity in a sponge larva. J Comp Physiol A 188(3):199–202

    Article  Google Scholar 

  • Leys SP, Mackie GO, Meech RW (1999) Impulse conduction in a sponge. J Exp Biol 202(9):1139–1150

    Google Scholar 

  • Leys SP, Meech RW (2006) Physiology of coordination in sponges. Can J Zool 84(2):288–306

    Article  Google Scholar 

  • Liebeskind BJ, Hillis DM, Zakon HH (2011) Evolution of sodium channels predates the origin of nervous systems in animals. PNAS 108(22):9154–9159

    Article  Google Scholar 

  • Lindgens D, Holstein TW, Technau U (2004) Hyzic, the Hydra homolog of the zic/odd-paired gene, is involved in the early specification of the sensory nematocytes. Development 131(1):191–201

    Article  Google Scholar 

  • Liu AG, McIlroy D, Brasier MD (2010) First evidence for locomotion in the Ediacara biota from the 565 Ma Mistaken Point Formation, Newfoundland. Geology 38(2):123–126

    Article  Google Scholar 

  • Lowe CJ, Wu M, Salic A, Evans L, Lander E, Strange-Thomann N, Gruber CE, Gerhart J, Kirschner M (2003) Anteroposterior patterning in hemichordates and the origins of the chordate nervous system. Cell 113(7):853–865

    Article  Google Scholar 

  • Maldonado M (2004) Choanoflagellates, choanocytes, and animal multicellularity. Invertebr Biol 123(1):1–22

  • Maldonado M (2006) The ecology of the sponge larva. Can J Zool 84(2):175–194

    Article  Google Scholar 

  • Maldonado M, Durfort M, McCarthy DA, Young CM (2003) The cellular basis of photobehavior in the tufted parenchymella larva of demosponges. Mar Biol 143(3):427–441

    Article  Google Scholar 

  • Marlow HQ, Srivastava M, Matus DQ, Rokhsar D, Martindale MQ (2009) Anatomy and development of the nervous system of Nematostella vectensis, an anthozoan cnidarian. Dev Neurobiol 69(4):235–254

    Article  Google Scholar 

  • Matus DQ, Pang K, Marlow H, Dunn CW, Thomsen GH, Martindale MQ (2006) Molecular evidence for deep evolutionary roots of bilateriality in animal development. PNAS 103(30):11195–11200

    Article  Google Scholar 

  • McCall G (2006) The Vendian (Ediacaran) in the geological record: enigmas in geology’s prelude to the Cambrian explosion. Earth-Sci Rev 77:1–229

    Article  Google Scholar 

  • McMenamin MAS (1986) The garden of Ediacara. PALAIOS 1(2):178–182

    Article  Google Scholar 

  • Miller H (1961) A generalization of Wald’s identity with applications to random walks. Ann Math Stat 32(2):549–560

    Article  MATH  Google Scholar 

  • Monk T, Green P, Paulin M (2014) Martingales and fixation probabilities of evolutionary graphs. Proc R Soc A 470(2165). doi:10.1098/rspa.2013.0730

  • Monk T, Paulin MG (2014) Predation and the origin of neurons. Brain Behav Evol 84:246–261

    Article  Google Scholar 

  • Moroz LL (2009) On the independent origins of complex brains and neurons. Brain Behav Evol 74(3):177–190

  • Moroz LL, Kocot KM, Citarella MR, Dosung S, Norekian TP, Povolotskaya IS, Grigorenko AP, Dailey C, Berezikov E, Buckley KM, Ptitsyn A, Reshetov D, Mukherjee K, Moroz TP, Bobkova Y, Yu F, Kapitonov VV, Jurka J, Bobkov YV, Swore JJ, Girardo DO, Fodor A, Gusev F, Sanford R, Bruders R, Kittler E, Mills CE, Rast JP, Derelle R, Solvyev VV, Kondrashov FA, Swalla BJ, Sweedler JV, Rogaev EI, Halanych KM, Kohn AB (2014) The ctenophore genome and the evolutionary origins of neural systems. Nature 510:109–114

    Article  Google Scholar 

  • Nakanishi N, Renfer E, Technau U, Rentzsch F (2012) Nervous systems of the sea anemone Nematostella vectensis are generated by ectoderm and endoderm and shaped by distinct mechanisms. Development 139(2):347–357

    Article  Google Scholar 

  • Narbonne GM (2005) The Ediacara biota: Neoproterozoic origin of animals and their ecosystems. Annu Rev Earth Planet Sci 33:421–442

    Article  Google Scholar 

  • Nickel M (2010) Evolutionary emergence of synaptic nervous systems: what can we learn from the non-synaptic, nerveless Porifera? Invertebr Biol 129(1):1–16

    Article  Google Scholar 

  • Niven JE (2005) Brain evolution: getting better all the time? Curr Biol 15(16):R624–R626

    Article  Google Scholar 

  • Niven JE, Anderson JC, Laughlin SB (2007) Fly photoreceptors demonstrate energy–information trade-offs in neural coding. PLoS Biol 5(4):e116

    Article  Google Scholar 

  • Niven JE, Laughlin SB (2008) Energy limitation as a selective pressure on the evolution of sensory systems. J Exp Biol 211(11):1792–1804

    Article  Google Scholar 

  • Northcutt RG (2012) Evolution of centralized nervous systems: two schools of evolutionary thought. PNAS 109:10626–10633

    Article  Google Scholar 

  • Parker AR (1998) Colour in Burgess Shale animals and the effect of light on evolution in the Cambrian. Proc R Soc B 265:967–972

    Article  Google Scholar 

  • Paterson JR, Garcia-Bellido DC, Lee MSY, Brock GA, Jago JB, Edgecombe GD (2011) Acute vision in the giant Cambrian predator Anomalocaris and the origin of compound eyes. Nature 480:237–240

    Article  Google Scholar 

  • Paulin MG (2005) Evolution of the cerebellum as a neuronal machine for Bayesian state estimation. J Neural Eng 2(3):S219–S234

    Article  Google Scholar 

  • Peterson KJ, Cotton JA, Gehling JG, Pisani D (2008) The Ediacaran emergence of bilaterians: congruence between the genetic and the geological fossil records. Philos Trans R Soc B 363:1435–1443

    Article  Google Scholar 

  • Peterson KJ, Lyons JB, Nowak KS, Takacs CM, Wargo MJ, McPeek MA (2004) Estimating metazoan divergence times with a molecular clock. PNAS 101(17):6536–6541

    Article  Google Scholar 

  • Peterson KJ, McPeek MA, Evans DAD (2005) Tempo and mode of early animal evolution: inferences from rocks, hox, and molecular clocks. Paleobiology 31(2–Suppl.):36–55

    Article  Google Scholar 

  • Pickard BG (1973) Action potentials in higher plants. Bot Rev 39(2):172–201

    Article  Google Scholar 

  • Reiczigel J (2003) Confidence intervals for the binomial parameter: some new considerations. Stat Med 22(4):611–621

    Article  Google Scholar 

  • Renard E, Vacelet J, Gazave E, Lapebie P, Borchiellini C, Ereskovsky AV (2009) Origin of the neuro-sensory system: new and expected insights from sponges. Integr Zool 4(3):294–308

    Article  Google Scholar 

  • Richards GS, Simionato E, Perron M, Adamska M, Vervoort M, Degnan BM (2008) Sponge genes provide new insight into the evolutionary origin of the neurogenic circuit. Curr Biol 18(15):1156–1161

    Article  Google Scholar 

  • Rivera AS, Ozturk N, Fahey B, Plachetzki DC, Degnan BM, Sancar A, Oakley TH (2012) Blue-light-receptive cryptochrome is expressed in a sponge eye lacking neurons and opsin. J Exp Biol 215(8):1278–1286

    Article  Google Scholar 

  • Ryan JF, Pang K, Schnitzler CE, Nguyen AD, Moreland RT, Simmons DK, Koch BJ, Francis WR, Havlak P, Smith SA, Putnam NH, Haddock SHD, Dunn CW, Wolfsberg TG, Mullikin JC, Martindale MQ, Baxevanis AD (2013) The genome of the ctenophore Mnemiopsis leidyi and its implications for cell type evolution. Science 342:1336–1343

    Article  Google Scholar 

  • Sakarya O, Armstrong KA, Adamska M, Adamski M, Wang IF, Tidor B, Degnan BM, Oakley TH, Kosik KS (2007) A post-synaptic scaffold at the origin of the animal kingdom. PLoS One 2(6):e506

    Article  Google Scholar 

  • Seilacher A (1999) Biomat-related lifestyles in the Precambrian. PALAIOS 14(1):86–93

    Article  Google Scholar 

  • Seilacher A, Buatois LA, Mangano MG (2005) Trace fossils in the Ediacaran–Cambrian transition: behavioral diversification, ecological turnover and environmental shift. Palaeogeogr Palaeoclimatol 227(4):323–356

    Article  Google Scholar 

  • Shen B, Dong L, Xiao S, Kowalewski M (2008) The Avalon explosion: evolution of Ediacara morphospace. Science 319:81–84

    Article  Google Scholar 

  • Slayman CL, Long WS, Gradmann D (1976) “Action potentials” in Neurospora crassa, a mycelial fungus. Biochim Biophys Acta 426:732–744

    Article  Google Scholar 

  • Sperling EA, Frieder CA, Raman AV, Girguis PR, Levin LA, Knoll AH (2013) Oxygen, ecology, and the Cambrian radiation of animals. PNAS 110(33):13446–13451

    Article  Google Scholar 

  • Stewart WJ, Cardenas GS, McHenry MJ (2013) Zebrafish larvae evade predators by sensing water flow. J Exp Biol 216(3):388–398

  • Tang F, Bengtson S, Wang Y, Wang XL, Yin CY (2011) Eoandromeda and the origin of Ctenophora. Evol Dev 13(5):408–414

    Article  Google Scholar 

  • Thieme H (2003) Mathematics in population biology. Princeton University Press, Princeton

    MATH  Google Scholar 

  • Tompkins-MacDonald GJ, Gallin WJ, Sakarya O, Degnan B, Leys SP, Boland LM (2009) Expression of a poriferan potassium channel: insights into the evolution of ion channels in metazoans. J Exp Biol 212(6):761–767

    Article  Google Scholar 

  • Trewavas A (2005) Plant intelligence. Naturwissenschaften 92(9):401–413

    Article  Google Scholar 

  • Tsutsui I, Ohkawa T, Nagai R, Kishimoto U (1987) Role of calcium ion in the excitability and electrogenic pump activity of the Chara corallina membrane: I. Effects of La3+, Verapamil, EGTA, W-7, and TFP on the action potential. J Membr Biol 96:65–73

    Article  Google Scholar 

  • Waggoner B (1998) Interpreting the earliest metazoan fossils: what can we learn? Am Zool 38(6):975–982

    Article  Google Scholar 

  • Wahab MAA, de Nys R, Whalan S (2011) Larval behaviour and settlement cues of a brooding coral reef sponge. Coral Reefs 30(2):451–460

    Article  Google Scholar 

  • Wald A (1944) On cumulative sums of random variables. Ann Math Stat 15(3):283–296

    Article  MATH  MathSciNet  Google Scholar 

  • Watanabe H, Fujisawa T, Holstein TW (2009) Cnidarians and the evolutionary origin of the nervous system. Dev Growth Differ 51(3):167–183

    Article  Google Scholar 

  • Whittle P (1964) Some general results in sequential analysis. Biometrika 51(1):123

    Article  MATH  MathSciNet  Google Scholar 

  • Wilensky U (1999) Netlogo. http://ccl.northwestern.edu/netlogo

  • Woollacott RM (1993) Structure and swimming behavior of the larva of Haliclona tubifera (Porifera: Demospongiae). J Morphol 218:301–321

    Article  Google Scholar 

  • Xiao S, Laflamme M (2008) On the eve of animal radiation: phylogeny, ecology and evolution of the Ediacara biota. Trends Ecol Evol 24(1):31–40

    Article  Google Scholar 

  • Xiao S, Yuan X, Knoll AH (2000) Eumetazoan fossils in terminal proterozoic phosphorites? PNAS 97(25):13684–13689

    Article  Google Scholar 

  • Zhang X, Shu D (2007) Soft anatomy of sunellid arthropods from the Chengjiang Lagerstatte, Lower Cambrian of southwest China. J Paleont 81(6):1412–1422

    Article  Google Scholar 

  • Zylberberg J, DeWeese MR (2011) How should prey animals respond to uncertain threats? Front Comput Neurosci 5:20–27

    Article  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Travis Monk.

Appendices

Appendix A: Derivation of \(p_{\text {fix}}\) for the idealised ecosystem

Our model ecosystem assumes that animals receive an independent and identically-distributed (i.i.d.) amount of energy on every timestep. Wald (1944) derived an identity that can be used to calculate not only absorption probabilities, but also the (conditional and marginal) distributions of the time to absorption, when the steps of a random walk are i.i.d.

Suppose that \(S_t\) is a sum of \(t\) i.i.d. random variables \(V_i\), where \(V_i \sim V\). Let the sum’s initial value \(S_0\) fall between two constant absorbing barriers \(b\) and \(a;\, b < S_0 < a\). As long as the sum stays between \(a\) and \(b\), we continue adding observations of \(V\) to the sum. When the sum hits either \(a\) or \(b\), it is said to ‘absorb.’ Let \(T\) represent the (finite) random number of summations that have been performed when the sum hits one of the two absorbing barriers.

Let \({\mathbb {M}}_{V}(h)\) be the moment generating function of \(V; \,{\mathbb {M}}_{V}(h) \equiv {\mathbb {E}}\left[ e^{Vh}\right] \), with \(h\) as a free variable. Wald’s identity is (Miller 1961; Whittle 1964) :

$$\begin{aligned} {\mathbb {E}}\left[ e^{S_Th}({\mathbb {M}}_{V}(h))^{-T}\right] =e^{S_0 h}. \end{aligned}$$
(5)

Lemma 2 in Wald (1944) states that if \({\mathbb {E}}\left[ V\right] \ne 0\) and \(\text {Var}[ V ] \ne 0\), then exactly one nonzero value \(h_0 \ne 0\) makes \({\mathbb {M}}_{V}(h_0) = 1\). In other words, \({\mathbb {M}}_{V}(h) - 1\) is convex and has two real roots, one nonzero, under very general assumptionsFootnote 1 (Barnett 1975). Evaluating Eq. 5 at \(h_0\) gives:

$$\begin{aligned} {\mathbb {E}}\left[ e^{S_Th_0}\right] = e^{S_0 h_0}. \end{aligned}$$
(6)

If the absorbing barriers 0 and \(V_{\text {rep}}\) are exactly reached upon absorption and not exceeded, then:

$$\begin{aligned} {\mathbb {E}}\left[ e^{S_T h_0}\right] = e^{a h_0} \Pr (S_T = a) + e^{b h_0} \Pr (S_T = b). \end{aligned}$$

Inserting \(\Pr (S_T = b) = 1 - \Pr (S_T = a)\) and rearranging, Eq. 6 becomes:

$$\begin{aligned} \Pr (S_T = a) = \frac{e^{S_0 h_0} - e^{b h_0}}{e^{a h_0} - e^{b h_0}}. \end{aligned}$$

For the animals considered in our model, insert \(S_0 = V_{\text {rep}}/2, a = V_{\text {rep}}\), and \(b = 0\) to obtain our claimed expression for \(p_{\text {rep}}\), Eq. 2.

Recall that \(h_0\) is the unique nonzero value that makes \({\mathbb {M}}_{V}(h_0) = 1\). When \({\mathbb {E}}\left[ V\right] \approx 0, h_0 \approx 0\), and we can approximate \(h_0\) very accurately by Taylor-expanding the moment generating function:

$$\begin{aligned} 1 = {\mathbb {M}}_{V}(h_0) \equiv {\mathbb {E}}\left[ e^{Vh_0}\right] = {\mathbb {E}}\left[ 1\right] + h_0 \cdot {\mathbb {E}}\left[ V\right] + h_0^2/2 \cdot {\mathbb {E}}\left[ V^2\right] + O(3). \end{aligned}$$

Neglecting higher-order terms and rearranging, we obtain Eq. 3:

$$\begin{aligned} h_0 \approx -2 {\mathbb {E}}\left[ V\right] / {\mathbb {E}}\left[ V^2\right] . \end{aligned}$$

We must stress that \({\mathbb {E}}\left[ V\right] \) must be very close to 0 for this approximation to be valid. A small error in our approximation of \(h_0\) can result in a large error for our approximation of \(p_{\text {rep}}\), so our approximation for \(h_0\) must be very accurate. More accurate approximations, if needed, can be obtained by numerical evaluation of \(h_0\) or by including more higher-order terms.

Finally we obtain the probability \(p_{\text {fix}}\) that a single decider will fix in a population of idiots.

When \({\mathbb {E}}\left[ V\right] \approx 0\), the equilibrium total (decider + idiot) animal population \(N\) will remain approximately constant throughout a decider’s invasion. Since \(N\) is approximately constant, we can approximate our model ecosystem (see Fig. 1) as a Moran process.

The Moran process is a birth-death process; each change in the population structure involves one individual reproducing (R) and one individual perishing (P). We can describe our idealised ecosystem as a birth-death process as follows. During the simulation, deciders (D) and idiots (I) reproduce and perish. Define a birth-death event to be the instant when a reproduction–perish pair [R, P], composed of one animal from each clade is completed (i.e. [R, P] = either [D, I] or [I, D]). The approximately constant population size ensures that the number of births and deaths will be approximately equal, so the reproduction and perish components of the [R, P] pairs will occur reasonably close together in time.

In the original Moran process, \(r\) is constant by definition, and \(p_{\text {fix}}\) is (Ewens 2004; Monk et al. 2014):

$$\begin{aligned} p_{\text {fix}}= (1 - r ^ {-1}) / (1 - r ^ {-N}). \end{aligned}$$

This expression for \(p_{\text {fix}}\) applies to our ecosystem if \(r\) is constant throughout a decider’s invasion. We now show that \(r\) indeed remains very nearly constant.

As in the Moran process, \(r\) is the ratio of the probabilities of the decider population size \(\delta _t\) increasing and decreasing by one:

$$\begin{aligned} r = \frac{\Pr (R = D, P = I| \delta _t)}{\Pr (R = I, P = D| \delta _t)} \approx \frac{\Pr (R = D| \delta _t) \Pr (P = I| \delta _t)}{\Pr (R = I| \delta _t) \Pr (P = D| \delta _t)}, \end{aligned}$$

since R and P are approximately conditionally independent, conditional on the decider population size. We may consider each of these probabilities as a ratio of reproduction or death rates \(q\):

(7)

where \(q_R^D\) is the rate that deciders reproduce, \(q_R^I\) is the rate that idiots reproduce, \(q_P^D\) is the rate that deciders perish, and \(q_P^I\) is the rate that idiots perish:

Here, \(p_{\text {rep}}^D\) is the probability that an individual decider will reproduce, and \({\mathbb {E}}[{T_D} | {R}]\) is the conditional expected time for an individual decider to absorb, given that it eventually reproduced. The other terms are defined analogously.

Since \({\mathbb {E}}\left[ V_t\right] \approx 0, {\mathbb {M}}_{V}(h)\) is well-approximated by a parabola in the neighbourhood \(h \approx 0\). This symmetry of \({\mathbb {M}}_{V}(h)\) and the symmetry of the absorbing barriers for individual animals (\(b = 0, S_0 = V_{\text {rep}}/2\), and \(a = V_{\text {rep}}\)) means that the conditional expected number of summations required to absorb, conditional on where absorption occurred, are equal. In other words, animals that start with energy \(V_{\text {rep}}/2\) and eventually reproduce require the same expected number of summations as those that start with energy \(V_{\text {rep}}/2\) and eventually starve.Footnote 2 This is shown in Fig. 5. Since \({\mathbb {E}}[{T_D} | {R}] = {\mathbb {E}}[{T_D} | {P}]\) and \({\mathbb {E}}[{T_I} | {R}] = {\mathbb {E}}[{T_I} | {P}]\), Eq. 7 simplifies to:

$$\begin{aligned} r = \frac{p_{\text {rep}}^D(1 - p_{\text {rep}}^I)}{p_{\text {rep}}^I(1 - p_{\text {rep}}^D)}. \end{aligned}$$
(8)

This expression for \(r\) is virtually independent of the decider population size \(\delta _t\), though not obviously so. \(p_{\text {rep}}^I\) and \(p_{\text {rep}}^D\) are dependent on \(\delta _t\) because, as an invasion progresses, the food item density changes. Before a decider is introduced, the food item density is \(\rho _I= (a+ m) / b\), where idiots break even on average. If a decider fixes, then the food item density has been driven to where deciders break even on average. This is shown in Fig. 5. Plotting \(r\) over a range of decider population sizes, we see that \(r\) remains approximately constant throughout a decider invasion (see Fig. 5).

Fig. 5
figure 5

\(r\) is virtually independent of the decider population size population size \(\delta _t\). Upper plot Conditional (triangles) and marginal (lines) expected times to absorption for deciders (solid line and filled triangles) and idiots (dashed line and empty triangles) as a function of \(\delta _t\). \({\mathbb {E}}[{T} | {R}] = {\mathbb {E}}[{T} | {P}]\), for both deciders and idiots, for all \(\delta _t\); all the triangles perfectly overlap. When an invasion begins, it takes idiots longer to reach absorption than deciders, but when an invasion is nearly complete, it takes deciders longer to absorb than idiots. Lower left \(p_{\text {rep}}\) as a function of \(\delta _t\). As the population size of advantaged deciders grows, food density \(\rho \) is driven down, so both deciders and idiots are less likely to reproduce as an invasion progresses. Lower right Plot of Eq. 8. \(r\) is approximately constant for all \(\delta _t\). The y-axis shows that the slope of the line is negligible. For all plots, at the beginning of a decider invasion, \({\mathbb {E}}\left[ V_t\right] \) = 3.5e-4, \({\mathbb {E}}\left[ V_t^2\right] = .2102, V_{\text {rep}}=500\), and \(N = 80\)

Since \(r\) is nearly constant with respect to \(\delta _t\), we can simplify Eq. 8 further by evaluating it at \(\delta _t = 1\). At \(\delta _t = 1, p_{\text {rep}}^I= 1/2\), so \(r\) simplifies to:

$$\begin{aligned} r = \left. \frac{p_{\text {rep}}^D}{1 - p_{\text {rep}}^D} \right| _{\rho = \rho _I}, \end{aligned}$$
(9)

where the evaluation at \(\rho =\rho _I\) means that we want to evaluate \(p_{\text {rep}}^D\) at food item density \(\rho _I= (a+ m) / b\).

Appendix B: Obtaining closed-form expressions for \(p_i\)

This Appendix defines models of the signals produced by food items and how a decider’s sensor spikes in response to those signals. It then shows how to calculate the \(p_i\) given these models.

Let \(r_k\) be the random distance between a decider’s sensor and the \(k\)th closest food item. Assume that food items produce signals \(\sigma _k\) that fall as distance cubed:

$$\begin{aligned} \sigma _k = 1/r_k^3. \end{aligned}$$

Let the sensor spike with an intensity \(I\) proportional to the additive strengths of the signals from the food items. The proportionality constant \(g\) is the gain of the sensor, and let \(n\) be the spontaneous firing rate:

$$\begin{aligned} I(\mathbf{{r}}) = g\sum _{k=1}^\infty \sigma _k + n. \end{aligned}$$

The sensor acts as a food item proximity detector; if food items are close to the sensor, then it experiences strong signals and spikes with a high intensity. The expected number of spikes \(\lambda \) that the sensor produces in time \(\varDelta t\) is:

$$\begin{aligned} \lambda (\mathbf{{r}}) = I(\mathbf{{r}}) \varDelta t. \end{aligned}$$

If the sensor spikes at least once in \(\varDelta t\), the decider strikes:

$$\begin{aligned} \Pr (A=1|\mathbf{{r}}) = 1 - e^{-\lambda (\mathbf{{r}})}. \end{aligned}$$

With this model of \(\Pr (A=1|\mathbf {r})\), we can calculate the probability that the decider strikes, given the random number of food items in the strike zone \(Z\) and given that the number of food items that exist inside the decider \(D\) is zero:Footnote 3

$$\begin{aligned} \Pr (A=1|Z,D=0)&= \int _0^\infty \Pr (A=1,\mathbf{{r}}|Z,D=0)d\mathbf{{r}} \\&= \int _0^\infty \Pr (A=1|\mathbf{{r}}) \Pr (\mathbf{{r}}|Z,D=0) d\mathbf{{r}} \\&= {\mathbb {E}}[{\Pr (A=1|\mathbf{{r}})} | {Z,D=0}] = {\mathbb {E}}[{1-e^{-\lambda }} | {Z,D=0}]. \end{aligned}$$

If the decider almost always strikes (i.e. if \(\Pr (A=1) \approx 1\)), then \(1-e^{-\lambda }\) is almost linear and Jensen’s inequality becomes an accurate approximation:

$$\begin{aligned} {\mathbb {E}}[{1-e^{-\lambda }} | {Z,D=0}] \approx 1 - \text {exp}\left( -{\mathbb {E}}[{\lambda } | {Z,D=0}] \right) . \end{aligned}$$

Inserting our expression for \(\lambda \):

$$\begin{aligned} {\mathbb {E}}[{1-e^{-\lambda }} | {Z,D=0}] \approx 1 - \text {exp}\left( -g\varDelta t \sum _{k=0}^\infty {\mathbb {E}}[{\sigma _k} | {Z,D=0}] - n\varDelta t \right) . \end{aligned}$$

The conditional expected signal of the \(k\)th closest food item given the number of food items in the zone and given that no food items exist inside the decider is:

$$\begin{aligned} {\mathbb {E}}[{\sigma _k} | {Z,D=0}]&= \int _0^\infty \sigma _k \Pr (r_k|Z,D=0)dr_k \\&= \int _0^\infty \sigma _k \frac{\Pr (Z,D=0|r_k)\Pr (r_k)}{\Pr (Z,D=0)}dr_k \\&= \int _0^\infty \sigma _k \frac{\Pr (Z|r_k,D=0)\Pr (D=0|r_k)\Pr (r_k)}{\Pr (Z)\Pr (D=0)} dr_k. \end{aligned}$$

All of the above probability density functions are known. Let \(\rho \) be the food item density, let the area of the decider be 1 (they are circular with radius \(\sqrt{1/\pi }\)) and let the area of the ‘strike zone’ be 1 (it is also circular with radius \(\sqrt{2/\pi }\), see Fig. 1). Then:

$$\begin{aligned} \Pr (Z)&= \text {Poisson}(\rho ); \qquad \Pr (D=0) = e^{-\rho }; \\ \Pr (r_k)&= 2(\pi \rho )^k r_k^{2k-1} \text {exp}\left( -\pi \rho r_k^2 \right) / (k-1)!; \\ \Pr (D=0|r_k)&= \left\{ \begin{aligned}&0, \\&\text {Binom}(0;k-1,1/\pi r_k^2) \\ \end{aligned} \right. \qquad \qquad \begin{aligned}&r_k \le \sqrt{1/\pi } \\&r_k > \sqrt{1/\pi }; \\ \end{aligned} \\ \Pr (Z|r_k,D=0)&= \left\{ \begin{aligned}&\text {Poisson}(2\rho - \rho \pi r_k^2),\\&\text {Binom}(Z;k-1,1/(\pi r_k^2 - 1)),\\ \end{aligned} \right. \qquad \begin{aligned}&Z \ge k \\&Z < k. \end{aligned} \end{aligned}$$

\(\Pr (A=1|Z,D=0)\) may now be numerically evaluated.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Monk, T., Paulin, M.G. & Green, P. Ecological constraints on the origin of neurones. J. Math. Biol. 71, 1299–1324 (2015). https://doi.org/10.1007/s00285-015-0862-7

Download citation

  • Received:

  • Revised:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00285-015-0862-7

Keywords

Mathematics Subject Classification

Navigation