14
LETTER doi:10.1038/nature13622 Dietary specializations and diversity in feeding ecology of the earliest stem mammals Pamela G. Gill 1 , Mark A. Purnell 2 , Nick Crumpton 1 {, Kate Robson Brown 3 , Neil J. Gostling 1 {, M. Stampanoni 4,5 & Emily J. Rayfield 6 The origin and radiation of mammals are key events in the history of life, with fossils placing the origin at 220 million years ago, in the Late Triassic period 1 . The earliest mammals, representing the first 50 million years of their evolution and including the most basal taxa, are widely considered to be generalized insectivores 1,2 . This implies that the first phase of the mammalian radiation—associated with the appearance in the fossil record of important innovations such as heterodont dentition, diphyodonty and the dentary–squamosal jaw joint 1,3 —was decoupled from ecomorphological diversification 2,4 . Finds of exceptionally complete specimens of later Mesozoic mammals have revealed greater ecomorphological diversity than previously suspected, including adaptations for swimming, burrowing, digging and even gliding 2,5,6 , but such well-preserved fossils of earlier mammals do not exist 1 , and robust analysis of their ecomorphological diversity has previously been lacking. Here we present the results of an inte- grated analysis, using synchrotron X-ray tomography and analyses of biomechanics, finite element models and tooth microwear tex- tures. We find significant differences in function and dietary ecology between two of the earliest mammaliaform taxa, Morganucodon and Kuehneotherium—taxa that are central to the debate on mammalian evolution. Morganucodon possessed comparatively more forceful and robust jaws and consumed ‘harder’ prey, comparable to extant small- bodied mammals that eat considerable amounts of coleopterans. Kuehneotherium ingested a diet comparable to extant mixed fee- ders and specialists on ‘soft’ prey such as lepidopterans. Our results reveal previously hidden trophic specialization at the base of the mam- malian radiation; hence even the earliest mammaliaforms were begin- ning to diversify—morphologically, functionally and ecologically. In contrast to the prevailing view 2,4 , this pattern suggests that lineage splitting during the earliest stages of mammalian evolution was asso- ciated with ecomorphological specialization and niche partitioning. Recently, much progress has been made in understanding the pattern and timing of the radiation of mammals 7–9 , revealing successive waves of taxonomic and ecomorphological diversification in Middle–Late Jurassic to Palaeogene stem clades and crown groups 2,10,11 . However, understand- ing of early mammaliaforms and the initial radiation of mammals has lagged behind. Here we address this problem by testing the hypothe- sis that two of the earliest and most basal mammaliaforms were eco- morphologically distinct. Morganucodon watsoni 12 and Kuehneotherium praecursoris 13 are central to the debate on mammalian origins and are of fundamental phylogenetic importance (Extended Data Fig. 1). Morga- nucodon is one of the earliest (Late Triassic to Early Jurassic) and best- known Mesozoic mammals, with a global distribution; Kuehneotherium is of a similar age and size 1,12,13 . Both taxa are thought to be generalized insectivores 1 and co-existed (see Supplementary Information for discus- sion of sympatry) on a small landmass present during the Early Jurassic marine transgression (Hettangian–Early Sinemurian, about 200 Myr ago), in what is now Glamorgan, South Wales, UK 1,12 (Extended Data Fig. 2). In addition to the apomorphic mammalian jaw joint, both taxa retain the plesiomorphic articular–quadrate jaw joint, as indicated by a well-developed postdentary trough (Fig. 1a, b), thus indicating that the postdentary bones still functioned as part of the jaw joint, rather than being incorporated into a definitive mammalian middle ear as in modern mammals 1,2,12–14 (sensu ref. 15). Curiously, Kuehneotherium possesses advanced molars, with cusps arranged in an obtuse-triangle pattern 13,16 (Extended Data Fig. 3b). We tested hypotheses of functional and dietary specialization in these early mammaliaforms by generating digital mandibular reconstructions, and applying a suite of techniques: classical mechanics, finite element modelling and quantitative textural analysis of tooth microwear. The mandible is a good choice for study of feeding adaptations as it is pri- marily adapted for biting, and is not constrained by sensory systems such as eye or brain size 17 . Our null hypothesis was that functional perfor- mance did not differ between the two taxa. Applying classical mechanics, we calculated the mechanical advan- tage for mid-molar, premolar and canine bites, reflecting the efficiency of the jaw system at transmitting force from the adductor muscles to the bite point. This revealed that Morganucodon has a notably larger mechan- ical advantage than Kuehneotherium (almost 50% greater during mid- molar biting) (Table 1), indicating that the mandible of Morganucodon had the potential to generate much larger bite forces than Kuehneotherium, and implying that Kuehneotherium bites were potentially faster but less forceful. We also determined jaw strength in bending and torsion dur- ing biting, treating the mandibular corpus as a beam 18 . The pattern of bending strength reveals a very different profile between the two taxa (Fig. 1c, d). Morganucodon shows peak resistance to bending at the rear of the tooth row as might be expected, as this region serves as a structural linkage between the tooth row and posterior functional elements of the jaw, such as the jaw joint and muscles 19 . However, Kuehneotherium shows peak resistance in the region of the anterior molars. Resistance to tor- sion (J) shows similar patterns (Fig. 1e, f). This different biomechanical profile in Kuehneotherium may reflect the importance of resisting bend- ing in the central tooth row, to maintain the sharp bladed triangulated molars in precise occlusion 16 . Finite element analysis allowed us to calculate stress, strain and defor- mation to assess the mechanical behaviour of the jaws 20 . This analysis can provide informative comparative data in the absence of known input parameters 17 and as such the two taxa were loaded with equal adductor muscle forces and constrained at the jaw joint and bite points (Extended Data Fig. 4). Finite element analysis shows that, during a simulated bite, despite similar length and surface area, the dentary of Kuehneotherium experiences greater maximum von Mises stress and maximum principal strain than Morganucodon, regardless of bite position, and higher reac- tion forces at the jaw joint, despite generating consistently less bite reac- tion force (Fig. 1g, h and Table 1). Kuehneotherium does not possess a robust condyle as in Morganucodon (Fig. 1a, b), further reducing its abil- ity to withstand high reaction forces at the jaw joint. We tested whether Morganucodon or Kuehneotherium could generate enough bite force to 1 School of Earth Sciences, University of Bristol, Wills Memorial Building, Queens Road, Bristol BS8 1RJ, UK. 2 Department of Geology, University of Leicester, University Road, Leicester LE1 7RH, UK. 3 Department of Archaeology and Anthropology, University of Bristol, Woodland Road, Bristol BS8 1UU, UK. 4 Swiss Light Source, Paul Scherrer Institute, CH-5232 Villigen, Switzerland. 5 Institute for Biomedical Engineering, University and ETH Zu ¨ rich, Gloriastrasse 35, CH-8092 Zu ¨ rich, Switzerland. 6 School of Earth Sciences, University of Bristol, Life Sciences Building, 24 Tyndall Avenue, Bristol BS8 1TQ, UK. {Present addresses: Natural History Museum, Cromwell Road, London SW7 5BD, UK (N.C.); Ocean and Earth Science, National Oceanography Centre, University of Southampton, Southampton SO14 3ZH, UK (N.J.G.). 21 AUGUST 2014 | VOL 512 | NATURE | 303 Macmillan Publishers Limited. All rights reserved ©2014

Dietary specializations and diversity in feeding ecology of the earliest stem mammals

  • Upload
    emily-j

  • View
    212

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Dietary specializations and diversity in feeding ecology of the earliest stem mammals

LETTERdoi:10.1038/nature13622

Dietary specializations and diversity in feedingecology of the earliest stem mammalsPamela G. Gill1, Mark A. Purnell2, Nick Crumpton1{, Kate Robson Brown3, Neil J. Gostling1{, M. Stampanoni4,5 & Emily J. Rayfield6

The origin and radiation of mammals are key events in the historyof life, with fossils placing the origin at 220 million years ago, in theLate Triassic period1. The earliest mammals, representing the first50 million years of their evolution and including the most basal taxa,are widely considered to be generalized insectivores1,2. This impliesthat the first phase of the mammalian radiation—associated withthe appearance in the fossil record of important innovations such asheterodont dentition, diphyodonty and the dentary–squamosal jawjoint1,3—was decoupled from ecomorphological diversification2,4. Findsof exceptionally complete specimens of later Mesozoic mammals haverevealed greater ecomorphological diversity than previously suspected,including adaptations for swimming, burrowing, digging and evengliding2,5,6, but such well-preserved fossils of earlier mammals donot exist1, and robust analysis of their ecomorphological diversityhas previously been lacking. Here we present the results of an inte-grated analysis, using synchrotron X-ray tomography and analysesof biomechanics, finite element models and tooth microwear tex-tures. We find significant differences in function and dietary ecologybetween two of the earliest mammaliaform taxa, Morganucodon andKuehneotherium—taxa that are central to the debate on mammalianevolution. Morganucodon possessed comparatively more forceful androbust jaws and consumed ‘harder’ prey, comparable to extant small-bodied mammals that eat considerable amounts of coleopterans.Kuehneotherium ingested a diet comparable to extant mixed fee-ders and specialists on ‘soft’ prey such as lepidopterans. Our resultsreveal previously hidden trophic specialization at the base of the mam-malian radiation; hence even the earliest mammaliaforms were begin-ning to diversify—morphologically, functionally and ecologically. Incontrast to the prevailing view2,4, this pattern suggests that lineagesplitting during the earliest stages of mammalian evolution was asso-ciated with ecomorphological specialization and niche partitioning.

Recently, much progress has been made in understanding the patternand timing of the radiation of mammals7–9, revealing successive waves oftaxonomic and ecomorphological diversification in Middle–Late Jurassicto Palaeogene stem clades and crown groups2,10,11. However, understand-ing of early mammaliaforms and the initial radiation of mammals haslagged behind. Here we address this problem by testing the hypothe-sis that two of the earliest and most basal mammaliaforms were eco-morphologically distinct. Morganucodon watsoni12 and Kuehneotheriumpraecursoris13 are central to the debate on mammalian origins and are offundamental phylogenetic importance (Extended Data Fig. 1). Morga-nucodon is one of the earliest (Late Triassic to Early Jurassic) and best-known Mesozoic mammals, with a global distribution; Kuehneotheriumis of a similar age and size1,12,13. Both taxa are thought to be generalizedinsectivores1 and co-existed (see Supplementary Information for discus-sion of sympatry) on a small landmass present during the Early Jurassicmarine transgression (Hettangian–Early Sinemurian, about 200 Myr ago),in what is now Glamorgan, South Wales, UK1,12 (Extended Data Fig. 2).In addition to the apomorphic mammalian jaw joint, both taxa retain the

plesiomorphic articular–quadrate jaw joint, as indicated by a well-developedpostdentary trough (Fig. 1a, b), thus indicating that the postdentary bonesstill functioned as part of the jaw joint, rather than being incorporatedinto a definitive mammalian middle ear as in modern mammals1,2,12–14

(sensu ref. 15). Curiously, Kuehneotherium possesses advanced molars,with cusps arranged in an obtuse-triangle pattern13,16 (Extended DataFig. 3b).

We tested hypotheses of functional and dietary specialization in theseearly mammaliaforms by generating digital mandibular reconstructions,and applying a suite of techniques: classical mechanics, finite elementmodelling and quantitative textural analysis of tooth microwear. Themandible is a good choice for study of feeding adaptations as it is pri-marily adapted for biting, and is not constrained by sensory systems suchas eye or brain size17. Our null hypothesis was that functional perfor-mance did not differ between the two taxa.

Applying classical mechanics, we calculated the mechanical advan-tage for mid-molar, premolar and canine bites, reflecting the efficiencyof the jaw system at transmitting force from the adductor muscles to thebite point. This revealed that Morganucodon has a notably larger mechan-ical advantage than Kuehneotherium (almost 50% greater during mid-molar biting) (Table 1), indicating that the mandible of Morganucodonhad the potential to generate much larger bite forces than Kuehneotherium,and implying that Kuehneotherium bites were potentially faster but lessforceful. We also determined jaw strength in bending and torsion dur-ing biting, treating the mandibular corpus as a beam18. The pattern ofbending strength reveals a very different profile between the two taxa(Fig. 1c, d). Morganucodon shows peak resistance to bending at the rearof the tooth row as might be expected, as this region serves as a structurallinkage between the tooth row and posterior functional elements of thejaw, such as the jaw joint and muscles19. However, Kuehneotherium showspeak resistance in the region of the anterior molars. Resistance to tor-sion (J) shows similar patterns (Fig. 1e, f). This different biomechanicalprofile in Kuehneotherium may reflect the importance of resisting bend-ing in the central tooth row, to maintain the sharp bladed triangulatedmolars in precise occlusion16.

Finite element analysis allowed us to calculate stress, strain and defor-mation to assess the mechanical behaviour of the jaws20. This analysiscan provide informative comparative data in the absence of known inputparameters17 and as such the two taxa were loaded with equal adductormuscle forces and constrained at the jaw joint and bite points (ExtendedData Fig. 4). Finite element analysis shows that, during a simulated bite,despite similar length and surface area, the dentary of Kuehneotheriumexperiences greater maximum von Mises stress and maximum principalstrain than Morganucodon, regardless of bite position, and higher reac-tion forces at the jaw joint, despite generating consistently less bite reac-tion force (Fig. 1g, h and Table 1). Kuehneotherium does not possess arobust condyle as in Morganucodon (Fig. 1a, b), further reducing its abil-ity to withstand high reaction forces at the jaw joint. We tested whetherMorganucodon or Kuehneotherium could generate enough bite force to

1School of Earth Sciences, University of Bristol, Wills Memorial Building, Queens Road, Bristol BS8 1RJ, UK. 2Department of Geology, University of Leicester, University Road, Leicester LE1 7RH, UK.3Department of Archaeology and Anthropology, University of Bristol, Woodland Road, Bristol BS8 1UU, UK. 4Swiss Light Source, Paul Scherrer Institute, CH-5232 Villigen, Switzerland. 5Institute forBiomedical Engineering, University and ETH Zurich, Gloriastrasse 35, CH-8092 Zurich, Switzerland. 6School of Earth Sciences, University of Bristol, Life Sciences Building, 24 Tyndall Avenue, Bristol BS81TQ, UK. {Present addresses: Natural History Museum, Cromwell Road, London SW7 5BD, UK (N.C.); Ocean and Earth Science, National Oceanography Centre, University of Southampton, SouthamptonSO14 3ZH, UK (N.J.G.).

2 1 A U G U S T 2 0 1 4 | V O L 5 1 2 | N A T U R E | 3 0 3

Macmillan Publishers Limited. All rights reserved©2014

Page 2: Dietary specializations and diversity in feeding ecology of the earliest stem mammals

pierce ‘hard’ insect cuticle (where ‘hard’ and ‘soft’ refer to the ease withwhich prey is pierced and chewed21). Estimation of bite force can cir-cumscribe the range of potential prey, providing a measure of feedingperformance and ecological partitioning22,23. A variety of insect prey wasavailable at the time: the Glamorgan fissures have yielded beetle remains24,and soft-bodied insects, such as scorpion flies, were well established inthe Early Jurassic25. (See Supplementary Information for discussion ofpotential prey.) A bite of 2 N is required to pierce the cuticle of a ‘hard’insect (for example, a beetle) of appropriate prey size for Morganucodonor Kuehneotherium26,27. For Morganucodon, a simulated 2 N bite at mid-molar m2 (see Methods) did not generate excessive stress in the jaw(maximum 54 MPa) (Fig. 1g). For Kuehneotherium, increasing muscleloadings (keeping the ratio of muscle recruitment intact), to simulate abite of 2 N at mid-molar m3 (Fig. 1i), produced higher reaction forcesat the dentary condyle (5.45 N compared with 2.38 N for Morganucodon),and maximum von Mises stress values up to 134 MPa, which is 2.5 timesthat of Morganucodon and close to the value of tensile stress failure forbone28. This suggests that Kuehneotherium was probably incapable ofprocessing ‘hard’ cuticle, and further illustrates differences in thebiomechanical performance of the jaws. Comparative biomechanicaldata therefore point to morphofunctional and dietary specialization inthese two taxa.

The hypothesis that Morganucodon and Kuehneotherium consumeddifferent prey was independently tested by comparing their tooth microwear

textures with those of extant insectivores with known dietary preferences(specimens listed in Extended Data Table 1). Recent work on insecti-vorous bats has shown that microwear textural analysis based on three-dimensional roughness parameters discriminates between insectivorespecies that consume different proportions of ‘hard’ prey (such as beetles)and ‘soft’ prey (such as moths)29. Bats provide a useful comparative dataset for our work because of their well-studied dietary differences and sim-ilarity in size to Morganucodon and Kuehneotherium. We compared thefossil taxa with four species of bats: Plecotus auritus (brown long-earedbat; a specialist on ‘soft’ insects); Pipistrellus pipistrellus (common pip-istrelle) and Pipistrellus pygmaeus (soprano pipistrelle) (more mixed diet,both specialize on Diptera (flies), but P. pipistrellus consumes insects witha wider range of cuticle ‘hardness’ and more ‘hard’ prey than P. pygmaeus);and Rhinolophus ferrumequinum (greater horseshoe bat; mixed diet,but including more beetles—prey that is among the ‘hardest’ of insects)(see Extended Data Table 2 for dietary details). In the bats, nine roughnessparameters differ significantly between species29, and principal compo-nent analysis (PCA) of these parameters (Fig. 2) separates bats accord-ing to dietary preferences in a space defined by principal component axes1 and 2 (together accounting for 88.3% of variance), with axis 1 stronglycorrelated with dietary preferences (rs 5 0.81, P , 0.0001). Increasinglynegative values indicate higher proportions of ‘hard’ prey, while increas-ing positive values indicate increasing proportions of ‘soft’ prey29.

Projecting data for Kuehneotherium and Morganucodon onto the axesresulting from the analysis of bats produces clear separation of the twotaxa. Morganucodon has negative values for principal component 1 (PC1),overlapping and extending beyond values for R. ferrumequinum. Slightlyrougher textures in Morganucodon suggest that it consumed a higherproportion of ‘hard’ prey. Most Kuehneotherium specimens have pos-itive values for PC1, overlapping the range of the ‘soft’ insect specialistPl. auritus. Two specimens have negative PC1 values and plot into aspace defined by the mixed-feeding Pipistrellus. Thirteen roughness para-meters from Morganucodon and Kuehneotherium are correlated withthe bat dietary axis (PC1; Extended Data Tables 3–5), including nine ofthe ten parameters that in bats are correlated with diet29, and values forPC1 differ significantly between the two fossil species (F 5 5.67; d.f. 5 6,29; P 5 0.0005). Pairwise tests (Tukey’s honestly significant difference;P , 0.05) indicate that microwear textures in Morganucodon and Kueh-neotherium differ from one another, yet Morganucodon does not differfrom bats with mixed or ‘harder’ diets, and Kuehneotherium does notdiffer from the ‘soft’ insect specialist and mixed feeders. Kuehneotheriumspecimens from different fissure localities do not differ from one another(see Supplementary Information for specimen and fissure details). ThatKuehneotherium and Morganucodon are so clearly separated by applicationof PCA based on extant bats with different diets provides powerful evi-dence that the two fossil taxa had diets that differed significantly in terms

m2 p4

0

CondylePostdentary trough

c m3 p6 c

Symphysis

0.8

0.4

1.5

1.0

0.5

m6 m1 p1 c0

a b

c d

e

g

i

h

f

Maximum stress: 54 MPa

2.38 N2.00 N

von Mises stress

054MPa

Maximum stress: 134 MPa

5.45 N 2.00 N

Maximum stress: 77 MPa

3.12 N 1.14 N

5 mm

0

0.8

0.4

m4 m1 p1 c

1.5

1.0

0.5

0

m1 p1 c m6 m1 p1 cm4

Figure 1 | Digital reconstructions and biomechanical analyses ofMorganucodon and Kuehneotherium jaws. a, b, Reconstructed leftmandibles, medial view. Missing portions in grey. a, M. watsoni.b, K. praecursoris; dentition reconstructed from ref. 30. c, d, Section moduli(dorsoventral Zx (blue) and mediolateral Zy (red), in cubic millimetres) for(c) Morganucodon and (d) Kuehneotherium. e, f, Polar moment of inertia (J)for (e) Morganucodon and (f) Kuehneotherium. g–i, Finite element vonMises stress contour plots for a unilateral molar bite (m2 for Morganucodonand m3 for Kuehneotherium) with closed jaw: g, Morganucodon andh, Kuehneotherium, with standard muscle loading; i, Kuehneotherium withmuscle loading multiplied to give a bite reaction force of 2 N. Stress legend andcontour plot scale apply to g–i. Single vertical arrows indicate tooth constraints;three grouped arrows indicate jaw joint constraints. (See SupplementaryInformation for links to scan data and finite element model images.)

Table 1 | Biomechanical analysis resultsMorganucodon Kuehneotherium

Mechanical advantage (three-dimensional)Canine bite 0.51 0.31Premolar bite 0.42 0.24Molar bite 0.35 0.18Maximum von Mises stress (MPa)Standard loading canine bite 66 97Standard loading premolar bite 61 83Standard loading molar bite 54 77Maximum principal strain (microstrain)Standard loading canine bite 3,840 5,800Standard loading premolar bite 3,540 5,020Standard loading molar bite 3,100 4,440Reaction forces (N) Jaw joint Bite Jaw joint BiteStandard loading canine bite 2.96 1.31 3.45 0.66Standard loading premolar bite 2.69 1.62 3.30 0.87Standard loading molar bite 2.38 2.00 3.12 1.14

Comparative mechanical advantage, maximum von Mises stress values (in megapascals), maximumprincipal strain values and reaction forces (in newtons) for Morganucodon and Kuehneotheriumdentaries. See Methods section for an explanation of the standard loading for muscles. These loadsachieve a 2 N reaction bite force at the molar (m2) of Morganucodon, sufficient to pierce insect cuticle.

RESEARCH LETTER

3 0 4 | N A T U R E | V O L 5 1 2 | 2 1 A U G U S T 2 0 1 4

Macmillan Publishers Limited. All rights reserved©2014

Page 3: Dietary specializations and diversity in feeding ecology of the earliest stem mammals

of prey ‘hardness’, and provides independent validation of distinctivemechanical behaviour and function revealed through our standard beamanalysis and finite element modelling.

In summary, our analyses reveal previously hidden trophic diversityand niche partitioning at the base of the mammalian radiation, support-ing a hypothesis of coupled lineage splitting and ecomorphological adap-tation of the skull and jaws, even during the earliest stages of mammalianevolution. Our approach, combining biomechanical analyses with toothmicrotextural validation of dietary differences, does not require excep-tionally preserved specimens, and is applicable to fragmentary fossilremains. As such, it has the potential to provide direct evidence of eco-morphology and adaptation through a range of vertebrate radiations,using the most commonly preserved fossil elements: teeth and jaws.

Online Content Methods, along with any additional Extended Data display itemsandSourceData, are available in the online version of the paper; references uniqueto these sections appear only in the online paper.

Received 8 April; accepted 27 June 2014.

1. Kielan-Jaworowska, Z., Cifelli, R. & Luo, Z.-X. Mammals from the Age of Dinosaurs.Origins, Evolution, and Structure (Columbia Univ. Press, 2004).

2. Luo, Z.-X. Transformation and diversification in early mammal evolution. Nature450, 1011–1019 (2007).

3. Kemp, T. S. The Origin and Evolution of Mammals Ch. 4 (Oxford Univ. Press, 2005).4. Bromham, L., Phillips, M. J. & Penny, D. Growing up with dinosaurs: molecular

dates and the mammalian radiation. Trends Ecol. Evol. 14, 113–118 (1999).5. Ji, Q., Luo, Z.-X., Yuan, C. X. & Tabrum, A. R. A swimming mammaliaform from the

Middle Jurassic and ecomorphological diversification of early mammals. Science311, 1123–1127 (2006).

6. Meng, J., Hu, Y., Wang, Y., Wang, X. & Li, C. A Mesozoic gliding mammal fromnortheastern China. Nature 444, 889–893 (2006).

7. Meredith, R. W. et al. Impacts of the Cretaceous terrestrial revolution and KPgextinction on mammal diversification. Science 334, 521–524 (2011).

8. dos Reis, M., Inoue, J., Asher, R. J., Donoghue, P. C. J. & Yang, Z. Phylogenomicdatasets provide both precision and accuracy in estimating the timescale ofplacental mammal phylogeny. Proc. R. Soc. B 279, 3491–3500 (2012).

9. O’Leary,M. A. et al. The placental mammal ancestor and the post–K-Pg radiation ofplacentals. Science 339, 662–667 (2013).

10. Wilson, G. P. et al. Adaptive radiation of multituberculate mammals before theextinction of dinosaurs. Nature 483, 457–460 (2012).

11. Grossnickle, D. M. & Polly, P. D. Mammal disparity decreases during theCretaceous angiosperm radiation. Proc. R. Soc. B 280, 20132110 (2013).

12. Kermack, K. A., Mussett, F. & Rigney, H. W. Lower jaw of Morganucodon. Zool. J. Linn.Soc. 53, 87–175 (1973).

13. Kermack,D.M., Kermack,K.A.&Mussett, F. TheWelshpantothereKuehneotheriumpraecursoris. Zool. J. Linn. Soc. 47, 407–423 (1968).

14. Luo, Z.-X. Developmental patterns in Mesozoic evolution of mammal ears. Annu.Rev. Ecol. Evol. Syst. 42, 355–380 (2011).

15. Allin, E. F. & Hopson, J. A. in The Evolutionary Biology of Hearing (eds Webster, D. B.,Fay, R. R. & Popper, A. N.) 587–614 (Springer, 1992).

16. Crompton, W. A. & Jenkins, F. A. American Jurassic symmetrodonts and Rhaeticpantotheres. Science 155, 1006–1009 (1967).

17. Tseng, Z. J., Mcnitt-Gray, J. L., Flashner, H., Wang, X. & Enciso, R. Model sensitivityand use of the comparative finite element method in mammalian jaw mechanics:mandible performance in the gray wolf. PLoS ONE 6, e19171 (2011).

18. Therrien, F. Mandibular force profiles of extant carnivorans and implications forthe feeding behaviour of extinct predators. J. Zool. 267, 249–270 (2005).

19. Freeman, P. W. & Lemen, C. A. Simple predictors of bite force in bats: the good, thebetter and the better still. J. Zool. 282, 284–290 (2010).

20. Rayfield, E. J. Finite element analysis and understanding the biomechanics andevolution of living and fossil organisms. Annu. Rev. Earth Planet. Sci. 35, 541–576(2007).

21. Evans, A. R. & Sanson, G. D. Biomechanical properties of insects in relation toinsectivory: cuticle thickness as an indicator of insect ‘hardness’ and‘intractability’. Aust. J. Zool. 53, 9–19 (2005).

22. Dumont, E. R. & Herrel, A. The effects of gape angle and bite point on bite force inbats. J. Exp. Biol. 206, 2117–2123 (2003).

23. Aguirre, L. F., Herrel, A., Van Damme, R. & Matthysen, E. Ecomorphological analysisof trophic niche partitioning in a tropical savannah bat community. Proc. R. Soc.Lond. B 269, 1271–1278 (2002).

24. Gardiner, B. G. New Rhaetic and Liassic beetles. Palaeontology 1, 87–88 (1961).25. Grimaldi,D.&Engel,M.S.Evolution of the Insects469–472 (CambridgeUniv. Press,

2005).26. Freeman, P. W. & Lemen, C. A. Using scissors to quantify hardness of insects: do

bats select or size or hardness? J. Zool. 271, 469–476 (2007).27. Aguirre, L. F., Herrel, A., van Damme, R. & Matthysen, E. The implications of food

hardness for diet in bats. Funct. Ecol. 17, 201–212 (2003).28. Currey, J. D. Bones: Structure and Mechanics 58–60 (Princeton Univ. Press, 2002).29. Purnell, M. A., Crumpton, N., Gill, P. G. & Rayfield, E. J. Within-guild dietary

discrimination from 3-D textural analysis of tooth microwear in insectivorousmammals. J. Zool. 291, 249–257 (2013).

30. Gill, P. G. Kuehneotherium from the Mesozoic Fissure Fillings of South Wales. PhDthesis, Univ. Bristol (2004).

Supplementary Information is available in the online version of the paper.

Acknowledgements We thank G. Armstrong, R. Asher, E. Bernard, P. Brewer, J. Bright,I. Corfe, A. Currant, A. Gill, T. Goddard, C. Hintermueller, J. Hooker, G. Jones,S. Lautenschlager, M. Lowe, F. Marone, F. Marx, C. Palmer, M. Pound, M. Ruecklin andJ. Sibbick. This work was funded by Natural Environment Research Council grants NE/E010431/1 and NE/K01496X/1 to E.J.R. and P.G.G.; M.A.P. was supported by NE/G018189/1.Useof the Swiss LightSource, PaulScherrer Institut, was supportedby theEuropean Commission 6th Framework Programme (RII3-CT-2004-506008).

Author Contributions E.J.R., P.G.G. and M.A.P. designed the study and wrote the paper;P.G.G., E.J.R., N.J.G. and M.S. collected the synchrotron radiation X-ray tomographicmicroscopy data; K.R.B. and P.G.G. collected the micro-computed tomography scandata; P.G.G. created the reconstructions and digital models, and analysed thebiomechanical results; P.G.G. and E.J.R. interpreted the biomechanical results; P.G.G.preparedand acquiredspecimens for microwearanalysis,M.A.P. and N.C. collected themicrowear data, M.A.P. analysed and interpreted the microwear results.

Author Information Reprints and permissions information is available atwww.nature.com/reprints. The authors declare no competing financial interests.Readers are welcome to comment on the online version of the paper. Correspondenceand requests for materials should be addressed to P.G. ([email protected]) orE.R. ([email protected]).

Plecotusauritus

Pipistrelluspipistrellus

Rhinolophusferrumequinum

Morganucodon (P3)Kuehneotherium (P3)Kuehneotherium (P5)

Increasing proportion of ‘hard’ prey in diet

–6

–4

–2

0

2

4

PC

2 (4

0.1%

)

–8 –6 –4 –2 0 2 4

PC1 (48.2%)

a b

c d

e

2

0

–2

2121

12121515

30302626

1111

6

3232

1010

2424

1414

3131

2727

28282525

1313

9

1616

82222

2323

2020

3

1717

5

3434

4

7

1919

1818

2929

2

1

333336363535

21

1215

3026

11

6

32

10

24

14

31

27

2825

13

9

16

822

23

20

3

17

5

34

4

7

19

18

29

2

1

333635

PipistrellusPipistrelluspygmaeuspygmaeusPipistrelluspygmaeus

Figure 2 | Quantitative textural analysis of microwear in bats and fossilmammaliaforms. a–d, Scale-limited roughness surfaces of Morganucodon(a; specimen 34), R. ferrumequinum (b; specimen 1), Kuehneotherium(c; specimen 24) and Pl. auritus (d; specimen 12); 146mm 3 110mm; contourvertical scale in micrometres. e, PCA of International Organization forStandardization (ISO) roughness parameters from bats and mammaliaforms.PCA based on data for bats only, with Morganucodon (n 5 5) andKuehneotherium data (Pontalum 3 (P3) n 5 5; Pant 5 fissure (P5) n 5 6)projected onto the bat PCA axes. There are two anomalous specimens: a singleKuehneotherium specimen (29) has PC1 values similar to R. ferrumequinum,and one of the Morganucodon specimens (32) plots as an outlier to all otherteeth analysed.

LETTER RESEARCH

2 1 A U G U S T 2 0 1 4 | V O L 5 1 2 | N A T U R E | 3 0 5

Macmillan Publishers Limited. All rights reserved©2014

Page 4: Dietary specializations and diversity in feeding ecology of the earliest stem mammals

METHODSDigital models. Digital mandibular reconstructions (Fig. 1a, b) were generated bycombining synchrotron radiation X-ray tomographic microscopy or micro-computedtomography scans from similar sized, mature individuals of each taxon (ExtendedData Fig. 3 and Supplementary Information). Morganucodon mandibular specimenswere scanned using synchrotron radiation X-ray tomographic microscopy at theSwiss Light Source TOMCAT beamline. The material was scanned at 18 keV and apixel size of 1.85mm (NHMUK PV M85507) or 22 keV and 3.7mm (UMZC Eo.D.45and UMZC Eo.D.66) to match the size of the sample with the field of view of themicroscope. The Kuehneotherium mandibular specimens and molar teeth, pro-vided to illustrate the difference in occlusion, were scanned on a Bruker Skyscan1172 micro-computed tomography system in the Department of Archaeology andAnthropology at the University of Bristol. A lower resolution was adequate for thefinite element models (20mm), although one specimen (NHMUK PV M92779) wasalso scanned at 3.1mm for internal detail.

Processing was performed using Avizo (Visualization Sciences Group). For eachtaxon, the computed tomography data for each specimen were re-oriented andscaled slightly if required, then digitally merged and manually rendered to producethe reconstructions. As all specimens were similar sizes, very limited scaling wasrequired. The final size of the reconstructed jaw was based on scaling all specimensto the size of the most complete specimen. Any empty alveoli were digitally infilledand tooth crowns removed to leave all tooth surfaces at an equivalent level to thetooth neck. Validation studies have shown that edentate finite element models betterrepresent experimental strains than dentate models31, which obviates the need toreconstruct missing teeth in the original scan data. A two-dimensional surface meshgenerated in Avizo (Visualization Sciences Group) was exported to Hypermesh (partof the Hyperworks suite from Altair), where the digital jaws were converted intofinite element models. Two-dimensional mesh optimization was performed and athree-dimensional finite-element mesh generated of linear four-noded tetrahedral(C3D4) elements (115,213 for Morganucodon and 68,555 for Kuehneotherium).Four-noded tetrahedra may be stiffer than ten-noded tetrahedra and may slightlyunderestimate strain32 but our models are comparative, and tetrahedral elementsare useful for modelling complex and intricate morphologies. The dentary and toothroots were created as separate parts with shared mesh boundaries. In the case of fossilmaterial, where it is impossible to validate predictions of bite force with in vivoexperimental data, it is possible to compare the relative performance of finite ele-ment models if the models are properly scaled. Scaling to equal force:surface arearatio provides a comparison of stress–strength performance based solely on shape33.The linear dimensions (canine to condyle) of the mandibles are 17.5 mm and 20.0 mmfor Morganucodon and Kuehneotherium respectively. The surface area ratio ofMorganucodon:Kuehneotherium, calculated from Avizo, is 1.02:1; as there is veryslightly more coronoid process and incisors missing from the Kuehneotheriummodel, the models are taken to be equal in surface area for this analysis. In this caseit was therefore not necessary to scale the applied muscle forces, and the modelscould be directly compared using the same loadings. The models were convertedto SI base-unit linear dimensions (metres) when imported into Abaqus/CAE forfinite element analysis.

The dentary bone was assigned isotropic and homogenous material properties:Young’s modulus of 18 GPa and Poisson’s ratio of 0.3. Properties for the tooth rootswere as for dentine: Young’s modulus of 25 GPa and Poisson’s ratio of 0.3. It is pos-sible that the actual jaw material properties were not isotropic, but in the absence ofdata otherwise, we assume isotropy for the sake of this analysis. Likewise, we willnever know the elastic properties of Mesozoic mammaliaform jaws, yet recent nano-indentation studies on the jaws of rats, squirrels and guinea pigs have revealed arange of 10–30 GPa for the Young’s modulus of bone, and 15–25 GPa for incisordentine34; our chosen values encompass this range.

The orientations (lines of action) of the mammaliaform anterior and posteriortemporalis and superficial and deep masseter adductor muscles were reconstructed(Extended Data Fig. 4a, b). The medial pterygoid was omitted as, if present, it wassmall35. Adductor muscle orientation was deduced by estimating the position of thepoint of origination on the skull, coupled with study of the insertion areas on themandibular fossae of the fossil specimens. The detailed description of the skull ofMorganucodon36 was used as reference for the muscle origins for Morganucodon.There is no skull material for Kuehneotherium; however, the gross morphology ofthe lower jaw of the slightly older Brazilian eucynodont Brasilitherium37,38 is verysimilar to Kuehneotherium (P.G.G., unpublished observations), so the general skullshape of Brasilitherium was used, with caution, as a proxy for the position of theKuehneotherium muscle origins (Extended Data Fig. 4c, d). The temporalis has itsorigin on the parietal bones of the skull, with development of a sagittal crest in Mor-ganucodon, although this latter is not known for Kuehneotherium. The temporalisinserts both medially and laterally in the temporal fossae on either side of the coro-noid process, with the muscles divided into anterior and posterior vectors35. The ante-rior temporalis insertions on the lateral and medial sides of the jaws were represented

by a predominantly vertical vector and were coupled to a single point in space rep-resenting their origin on the skull. The posterior temporalis component was dealtwith by applying a posteriorly directed load, as from the posterodorsal portion ofthe coronoid process. The coronoid hook of Morganucodon12 and posterodorsalportion of the coronoid process in Kuehneotherium are missing in the reconstruc-tions, so a means to simulate the missing portions of the coronoid processes for themuscle attachment was devised. A multi-point rigid body was created, from thebroken edge of the coronoid process to the position of the coronoid hook, in effectcompleting the coronoid process and producing a point in space for the posteriortemporalis origin (Extended Data Fig. 4 insert). The masseter in mammals is dividedinto superficial and deep components, with their origins on the zygomatic arch35: thesuperficial masseter at the anterior end on the jugal and the deep masseter poster-iorly on the squamosal36. They insert laterally on the jaw behind the tooth row. InMorganucodon, the two mandibular fossae are distinct and that of the superficialmasseter is adjacent to the angular process of the jaw. There is no angular process inKuehneotherium, but the masseteric fossa is well developed. The lines of action ofmuscles defined here were used to calculate mechanical advantage, and for the finiteelement analysis.Classical mechanics. The digital models were used to calculate mechanical advan-tage and beam theory to measure strength in bending. Mechanical advantage, as ameasure of the efficiency of the jaw system to transfer input muscle force from theadductor musculature to the point of biting, is frequently used as a metric of jawfunction39 and correlates with prey choice and feeding ecology in organisms suchas fish40,41. Mechanical advantage is calculated here as the length of the in-lever (themoment arm of the adductor musculature) divided by the length of the out-lever(the moment arm of the bite: the distance from the jaw joint to the bite point). Themoment arms were calculated by taking scaled screen images of the jaws in Abaqus,both in lateral and dorsal orientations to calculate the in-lever arm for each of thefour muscles (anterior temporalis, posterior temporalis, deep masseter and super-ficial masseter). A central point was chosen within the muscle insertion area, in eachcase, to give the line of action of each muscle. The four muscle vectors were resolvedto give a single adductor muscle vector, and the in-lever arm calculated.

Beam theory has been applied to the mandibular corpus in a number of studies18,22,42

and is related to dietary specialization in small mammals such as bats43. A measureof strength in bending is estimated from the section modulus Z, calculated at spe-cific intervals along the jaw. The section modulus Z is the second moment of area(I) divided by the distance from the neutral axis to the outer edge, in the plane ofbending, so the orientation considered affects the value of the bending strength. Inthis case the section modulus was measured in the dorsoventral (Zx) and medio-lateral planes (Zy). We follow ref. 18 in measuring Z at interdental gaps along thetooth row to the canine, plus a further section just posterior to the ultimate molar.The reconstructed hemimandibles were digitally sliced perpendicular to the longaxis of the mandible to produce cross-sectional images at the interdental gaps. Eachcross-sectional image was loaded into ImageJ44, and Zx and Zy calculated using theplugin MomentMacro. We measured all interdental gaps, with postcanine num-bers of 8 and 12 for Morganucodon and Kuehneotherium respectively. The polarmoment of inertia (J), the beam’s ability to resist torsion, is calculated from theaddition of the second moment of area, I, in the dorsoventral (Ix) and mediolateral(Iy) planes.Finite element analysis. Boundary constraints were applied to the condyle and atthree bite points: canine, final premolar and mid-molar. The anterior incisor bitepoint could not be included, as there are no specimens of this portion of the man-dible in Kuehneotherium. The ultimate premolar is the largest of the premolars inboth taxa. In Morganucodon, the largest (second) molar was chosen as the ‘mid molar’and in Kuehneotherium a mid-row (third) molar was used. Both single node45 anddistributed area (stiff beam elements) constraints46 have been used to estimate biteforces (discussion in ref. 17), but we used multipoint constraints, with master andslave nodes to minimize artificial stress concentrations47. There were approximately22 nodes constrained for each tooth and 30 nodes constrained at the jaw joint.Boundary conditions should never restrict deformations allowed by the representedenvironment47, so the bite points were appropriately constrained in four degrees offreedom (U1 5 U2 5 UR2 5 UR3 5 0), and the dentary condyle in four degrees offreedom (U1 5 U2 5 U3 5 UR3 5 0). (NB U1 is the mesiodistal axis, U2 is the dor-soventral axis and U3 is the axis along the length of the jaw; U refers to translationalmovement, UR refers to rotational movement.)

As the true muscle loadings for Morganucodon and Kuehneotherium are notknown, we used a comparative approach, assuming an equal muscle load appliedto each taxon48,49. Given the equivalence of surface area between the two models, thiswas appropriate33. For the initial models, the relative contribution of each muscle tooverall bite force was based on muscle ratios assigned to Morganucodon in ref. 35.The authors of this reference assign unit values of muscle forces to the jaw of Mor-ganucodon: anterior temporalis, 10; posterior temporalis, 8; superficial masseter, 8;deep masseter, 8. The values are arbitrary but reflect the relative proportions each

RESEARCH LETTER

Macmillan Publishers Limited. All rights reserved©2014

Page 5: Dietary specializations and diversity in feeding ecology of the earliest stem mammals

muscle contributes to bite force production. The actual loading forces we used wereas follows: anterior temporalis, 2 N; posterior temporalis, 1.6 N; superficial masseter,1.6 N; deep masseter, 1.6 N. We call this the standard loading regime. Contractingtogether at 100% activation, these muscles generate a bite reaction force of 2 N atthe mid-molar of the Morganucodon finite element model, sufficient to pierce insectcuticle (see below). The calculations in ref. 35 are based on a unilateral bite, but withadductors active on both sides, so allowance is made for the balancing side. BothMorganucodon and Kuehneotherium have a mobile symphysis12,30 and, as this cur-rent study is a comparative one of single lower jaws, it does not make assumptionsabout the forces on the balancing side, and the loads above are applied to the indi-vidual mandibular rami.

The reaction forces were queried in Abaqus at the bite points and condyle andthe maximum von Mises stress patterns recorded. Von Mises stress is calculated asit indicates regional deformation as a function of the three principal stresses s1, s2

and s3 (ref. 50). Maximum principal strain values were also recorded (Table 1). Muscleloadings were then manipulated to obtain a bite reaction force sufficient to pierceappropriately sized beetle carapace. Myotis bats are a similar size to Morganucodonand Kuehneotherium (skull length approximately 14 mm), and beetles in their stom-achs range in length from 1 to 10 mm, with a 10 mm beetle requiring 2–3 N of forceto pierce the insect27. This is corroborated by Myotis velifer bats recorded as havinga bite force of 2.2 N (ref. 26). The initial muscle loads in Morganucodon produced abite reaction force of 2 N at the molar bite, but in Kuehneotherium it was necessaryto increase the muscle loadings by 1.75 times to give a 2 N reaction force at the con-strained molar tooth.Microtextural analysis of tooth microwear. Recent work has shown that quant-itative microtextural analysis of tooth wear is a powerful tool for dietary discrim-ination and investigation of trophic resource exploitation in a range of extant andfossil vertebrates51–56. Our analysis compared the values for ISO 25178-2 areal tex-ture parameters57 for worn tooth surfaces in Morganucodon and Kuehneotheriumwith the results of analysis of the relationship between texture and diet in extantinsectivorous bats29. Three-dimensional microtextural analysis is entirely indepen-dent of our other functional and biomechanical analyses and thus provides effec-tive validation of our results.

Material used for microwear analysis is listed in Extended Data Table 1. Fossilspecimens were prepared at the School of Earth Sciences, University of Bristol, byimmersing dried blocks of matrix in hot water, with the addition of dilute hydro-gen peroxide only if required. The exception was two Pontalun 3 Kuehneotheriummolars, prepared at University College London in the 1970s, with added sodiumhexametaphosphate (Calgon) to aid dissolution of the matrix. No acetic acid was usedon any specimens. Bat specimens were all wild-found, acquired from UK sources(Extended Data Table 1) and assumed to be natural deaths. Specimens were fixedin either ethanol (University of Bristol and North Lancashire Bat Group) or 10%formalin solution (Veterinary Laboratory Agencies). Taxa were selected to includeinsectivores with well-constrained differences in their diets58,59. See Extended DataTable 2 and ref. 29 for details.

Our methods for capture of three-dimensional microwear data follow those devel-oped in ref. 60 (for full details see ref. 29). Three-dimensional surface data werecaptured from tooth wear facets (distal protoconid facet of m2 for bats, distobuccalwear facet of the main cusp of m2 for Morganucodon and a mid-row molar for Kueh-neotherium) using an Alicona Infinite Focus microscope G4b (IFM; software version2.1.2, field of view 145mm 3 110mm, lateral optical resolution 0.35–0.4mm, verticalresolution 20mm; lateral resolution factor for the IFM set at 0.3) (see ref. 29). Allthree-dimensional data were edited to delete dirt and dust particles from the sur-face (using Alicona IFM software) and exported as .sur files. All subsequent process-ing of data used SurfStand (version 5.0.0). Data were levelled, and a fifth-order robustpolynomial and a robust Gaussian wavelength filter (lc 5 0.025 mm) applied to removegross surface form and long-wavelength features of the tooth surface. This gener-ated a scale-limited roughness surface (Fig. 2) from which we derived ISO 25178-2standard roughness parameters57. Sample sizes used in this study are relatively small;however, as demonstrated in ref. 29, this does not prevent detection of dietary signalsthrough microtextural analysis. Data were explored using analysis of variance, cor-relations and principal components (on correlations; PCA). All statistical analysisof microtextural data used JMP 9. The results of Shapiro–Wilk tests indicated thatsome roughness parameters were non-normally distributed (P . 0.05), but for almostall parameters we were unable to reject the null hypothesis of normality for log-transformed data, so log-transformed data were used for analysis. Where homo-geneity of variance tests revealed evidence of unequal variances, Welch analysis ofvariance was used.

31. Marinescu, R., Daegling, D. J. & Rapoff, A. J. Finite-element modeling of theanthropoid mandible: the effects of altered boundary conditions. Anat. Rec. 283,300–309 (2005).

32. Bright, J. A. & Rayfield, E. J. The response of cranial biomechanical finiteelement models to variations in mesh density. Anat. Rec. 294, 610–620(2011).

33. Dumont, E. R., Grosse, I. R. & Slater, G. J. Requirements for comparing theperformance of finite element models of biological structures. J. Theor. Biol. 256,96–103 (2009).

34. Cox, P. G., Fagan, M. J., Rayfield, E. J. & Jeffery, N. Finite element modelling ofsquirrel, guinea pig and rat skulls: using geometric morphometrics to assesssensitivity. J. Anat. 219, 696–709 (2011).

35. Crompton, A. W. & Hylander, W. L. in The Ecology and Biology of Mammal-likeReptiles (eds Hotton, N., MacLean, P. D., Roth, J. J. & Roth, E. C.) 263–282(Smithsonian Institution, 1986).

36. Kermack, K. A., Mussett, F. & Rigney, H. W. The skull of Morganucodon. Zool. J. Linn.Soc. 71, 1–158 (1981).

37. Bonaparte, J. F., Martinelli, A. G., Schultz, C. L. & Rubert, R. The sister group ofmammals: small cynodonts from the Late Triassic of southern Brazil. Rev. Brasil.Paleont. 5, 5–27 (2003).

38. Bonaparte, J. F., Martinelli, A. G. & Schultz, C. L. New information on Brasilodon andBrasilitherium (Cynodontia, Probainognathia) from the Late Triassic, southernBrazil. Rev. Brasil. Paleont. 8, 25–46 (2005).

39. Hiiemae, K.W. & Houston, J. B. The structureand function of the jaw muscles in therat (Rattus norvegicus L.). Zool. J. Linn. Soc. 50, 75–99 (1971).

40. Wainwright, P. C., Bellwood, D. R., Westneat, M. W., Grubrich, J. R. & Hoey, A. S. Afunctional morphospace for the skull of labrid fishes: patterns of diversity in acomplex biomechanical system. Biol. J. Linn. Soc. 82, 1–25 (2004).

41. Bellwood, D. R., Wainwright, P. C., Fulton, C. J. & Hoey, A. S. Functional versatilitysupports coral reef biodiversity. Proc. R. Soc. B 273, 101–107 (2006).

42. Cuff, A. R. & Rayfield, E. J. Feeding mechanics in spinosaurid theropods and extantcrocodilians. PLoS ONE 8, e65295 (2013).

43. Dumont, E. R. & Nicolay, C. W. Cross-sectional geometry of the dentary in bats.Zoology 109, 66–74 (2007).

44. Rasband, W. S. ImageJ (US National Institutes of Health, Bethesda, Maryland,2012).

45. Davis, J. L., Santana, S. E., Dumont, E. R. & Grosse, I. R. Predicting bite force inmammals: two-dimensional versus three-dimensional lever models. J. Exp. Biol.213, 1844–1851 (2010).

46. McHenry, C. R., Wroe, S., Clausen, P. D., Moreno, K. & Cunningham, E.Supermodeled sabercat, predatory behavior in Smilodon fatalis revealed byhigh-resolution 3D computer simulation. Proc. Natl Acad. Sci. USA 104,16010–16015 (2007).

47. Donald, B. J. M. Practical Stress Analysis with Finite Elements Ch. 7 (Glasnevin,2007).

48. Christiansen, P. A dynamic model for the evolution of sabrecat predatory bitemechanics. Zool. J. Linn. Soc. 162, 220–242 (2011).

49. Tseng, Z. J. & Binder, W. J. Mandibular biomechanics of Crocuta crocuta, Canislupus, and the late Miocene Dinocrocuta gigantea (Carnivora, Mammalia). Zool.J. Linn. Soc. 158, 683–696 (2010).

50. Dumont, E. R., Piccirillo, J. & Grosse, I. R. Finite-element analysis of bitingbehavior and bone stress in the facial skeletons of bats. Anat. Rec. A 283, 319–330(2005).

51. Scott, R. S. et al. Dental microwear texture analysis shows within-species dietvariability in fossil hominins. Nature 436, 693–695 (2005).

52. Scott, R. S. et al. Dental microwear texture analysis: technical considerations.J. Hum. Evol. 51, 339–349 (2006).

53. Ungar, P.Dental microwear texture analysis of varswaterbovidsand Early Pliocenepaleoenvironments of Langebaanweg, Western Cape Province, South Africa.J. Mamm. Evol. 14, 163–181 (2007).

54. Ungar, P. S., Grine, F. E. & Teaford, M. F. Dental microwear and diet of the Plio-Pleistocene hominin Paranthropus boisei. PLoS ONE 3, e2044 (2008).

55. Purnell, M. A., Hart, P. J. B., Baines, D. C. & Bell, M. A. Quantitative analysis of dentalmicrowear in threespine stickleback: a new approach to analysis of trophicecology in aquatic vertebrates. J. Anim. Ecol. 75, 967–977 (2006).

56. Calandra, I., Schulz, E., Pinnow, M., Krohn, S. & Kaiser, T. M. Teasing apart thecontributions of hard dietary items on 3D dental microtextures in primates.J. Hum. Evol. 63, 85–98 (2012).

57. International Organization for Standardization. ISO 25178-2:2012. GeometricalProduct Specifications (GPS) – Surface Texture: Areal – Part 2: Terms, Definitionsand Surface Texture Parameters (International Organization for Standardization,2012).

58. Vaughan, N. The diets of British bats (Chiroptera). Mammal Rev. 27, 77–94 (1997).59. Barlow, K. E. The diets of two phonic types of the bat Pipistrellus pipistrellus in

Britain. J. Zool. 243, 597–609 (1997).60. Purnell, M. A., Seehausen, O. & Galis, F. Quantitative three-dimensional

microtextural analysis of tooth wear as a tool for dietary discrimination in fishes.J. R. Soc. Interface http://dx.doi.org/10.1098/rsif.2012.0140 (4 April 2012).

61. Luo, Z.-X., Kielan-Jaworowska, Z. & Cifelli, R. L. In quest for a phylogeny of Mesozoicmammals. Acta Palaeontol. Pol. 47, 1–78 (2002).

62. Evans, S. E. & Kermack, K. A. in In the Shadow of the Dinosaurs, Early MesozoicTetrapods (eds Fraser, N. C. & Sues, H.-D.) 271–283 (Cambridge Univ. Press,1994).

63. Gill, P. G. Resorption of premolars in the early mammal Kuehneotheriumpraecursoris. Arch. Oral Biol. 19, 327–328 (1974).

LETTER RESEARCH

Macmillan Publishers Limited. All rights reserved©2014

Page 6: Dietary specializations and diversity in feeding ecology of the earliest stem mammals

Extended Data Figure 1 | Phylogenetic relationships of major Mesozoicmammal lineages. Relative tree positions of Morganucodon andKuehneotherium in red. Based on figure 1 in ref. 61. The filled green circledenotes the node for the mammalian crown group.

RESEARCH LETTER

Macmillan Publishers Limited. All rights reserved©2014

Page 7: Dietary specializations and diversity in feeding ecology of the earliest stem mammals

Extended Data Figure 2 | Pontalun 3 fissure locality. a, Map to show locationof the Glamorgan quarries in South Wales, UK. The black square denotes thearea of the map shown in b. Map attribution: Jhamez84. CC-BY-SA-3.0 (http://

creativecommons.org/licenses/by-sa/3.0). b, Location of the Glamorganquarries yielding tetrapod remains, with white arrow marking Pontalun quarry.Carboniferous limestone upland areas in grey. Modified from ref. 62.

LETTER RESEARCH

Macmillan Publishers Limited. All rights reserved©2014

Page 8: Dietary specializations and diversity in feeding ecology of the earliest stem mammals

Extended Data Figure 3 | Molar form and specimens scanned to createthe digital reconstructions of the dentaries of Morganucodon andKuehneotherium. a, NHMUK PV M92538, an isolated M. watsoni right lowermolar (identified as m4) in lingual view. b, NHM NHMUK PV M9277, anisolated K. praecursoris right lower molar (mid-row) in lingual view; note thetriangulation of the cusp arrangement. Both molars digitally reconstructedfrom micro-computed tomography scans and reversed to fit with views of thedentaries. c–e, Specimens used for the digital reconstruction of M. watsoni:c, UMZC Eo.D.61 with m4 in situ; d, UMZC Eo.D.45 with p4, m1, m3 and m4in situ; e, NHMUK PV M85507 with i1–i4 (NB the only Glamorgan specimen

known with complete incisors in situ). f–i, Specimens used for the digitalreconstruction of K. praecursoris: f, NHMUK PV M19766 (paratype C865 inref. 13) with coronoid process and condylar region; g, NHMUK PV M19749(paratype C864 in ref. 13) postdentary trough region; h, UMZC Sy.97 withcomplete alveoli for m5–m6 and partial alveoli for m3–m4; i, NHMUK PVM92779 with alveoli for p1–m4 (U73 in ref. 63). All from Pontalun 3 fissure,except f and g from Pontalun 1 fissure, which ref. 30 assigned to the samehypodigm. All images show medial view. All are left dentaries, except c andd, which are reversed for ease of reference to the reconstructions in Fig. 1.

RESEARCH LETTER

Macmillan Publishers Limited. All rights reserved©2014

Page 9: Dietary specializations and diversity in feeding ecology of the earliest stem mammals

Extended Data Figure 4 | Static loaded finite element models to representthe jaw at the moment of biting. Right mandible models in lateral view of(a) Morganucodon and (b) Kuehneotherium to show muscle loading,constraints and bite points. Inset in a shows modelled rigid body used tosimulate missing coronoid process, for posterior temporalis loading. AT,anterior temporalis; PT, posterior temporalis; SM, superficial masseter; DM,deep masseter. Constraints indicated at the jaw joint and three individual bite

points at the mid molar (m2 in Morganucodon and m3 in Kuehneotherium),ultimate premolar (p4 in Morganucodon and p6 in Kuehneotherium) andcanine. The muscle origin positions are shown for (c) Morganucodon and(d) Kuehneotherium. Morganucodon skull reconstruction from ref. 36 andBrasilitherium skull, used as a proxy for the unknown Kuehneotherium skull,from ref. 37. The teeth have been removed for consistency with the mandiblemodels. See Methods for explanations.

LETTER RESEARCH

Macmillan Publishers Limited. All rights reserved©2014

Page 10: Dietary specializations and diversity in feeding ecology of the earliest stem mammals

Extended Data Table 1 | Specimens used for the microtextural analysis of tooth microwear

LEIUG, University of Leicester Geology collections; UOB, University of Bristol; VLA, Veterinary Laboratory Agencies; NLB, North Lancashire Bat Group.

RESEARCH LETTER

Macmillan Publishers Limited. All rights reserved©2014

Page 11: Dietary specializations and diversity in feeding ecology of the earliest stem mammals

Extended Data Table 2 | Trophic categorization and diets of British bat species used for validation of microtextural analysis of early mammalteeth

Table from ref. 29, modified from ref. 58 and references therein, and ref. 59.

LETTER RESEARCH

Macmillan Publishers Limited. All rights reserved©2014

Page 12: Dietary specializations and diversity in feeding ecology of the earliest stem mammals

Extended Data Table 3 | Correlations between roughness parameters from Morganucodon and Kuehneotherium teeth and principal com-ponent axes 1 and 2 derived from analysis of the nine roughness parameters that differ between bat species

*P , 0.05. See Extended Data Table 5 for definitions of parameters.

RESEARCH LETTER

Macmillan Publishers Limited. All rights reserved©2014

Page 13: Dietary specializations and diversity in feeding ecology of the earliest stem mammals

Extended Data Table 4 | Loadings (eigenvectors) for roughness para-metersontoprincipalcomponentaxes1and2for thePCAofbatspecies

From ref. 29.

LETTER RESEARCH

Macmillan Publishers Limited. All rights reserved©2014

Page 14: Dietary specializations and diversity in feeding ecology of the earliest stem mammals

Extended Data Table 5 | Short definitions and categorization of three-dimensional areal surface texture parameters

For further explanation, see ref. 29 and Supplementary Figs 1 and 2.

RESEARCH LETTER

Macmillan Publishers Limited. All rights reserved©2014