31.12.2013 Views

The primate cranial base: ontogeny, function and - Harvard University

The primate cranial base: ontogeny, function and - Harvard University

The primate cranial base: ontogeny, function and - Harvard University

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

YEARBOOK OF PHYSICAL ANTHROPOLOGY 43:117–169 (2000)<br />

<strong>The</strong> Primate Cranial Base:<br />

Ontogeny, Function, <strong>and</strong> Integration<br />

DANIEL E. LIEBERMAN, 1 CALLUM F. ROSS, 2 AND MATTHEW J. RAVOSA 3<br />

1 Department of Anthropology, George Washington <strong>University</strong>,<br />

Washington, DC 20052, <strong>and</strong> Human Origins Program, National Museum<br />

of Natural History, Smithsonian Institution, Washington, DC 20560<br />

2 Department of Anatomical Sciences, Health Sciences Center, SUNY at<br />

Stony Brook, Stony Brook, NY 11794-8081<br />

3 Department of Cell <strong>and</strong> Molecular Biology, Northwestern <strong>University</strong><br />

Medical School, Chicago, Illinois 60611-3008, <strong>and</strong> Mammals Division,<br />

Department of Zoology, Field Museum of Natural History,<br />

Chicago, Illinois 60605-2496<br />

<strong>cranial</strong> <strong>base</strong>; basicranium; chondrocranium; pri-<br />

KEY WORDS<br />

mates; humans<br />

ABSTRACT Underst<strong>and</strong>ing the complexities of <strong>cranial</strong> <strong>base</strong> development,<br />

<strong>function</strong>, <strong>and</strong> architecture is important for testing hypotheses about<br />

many aspects of craniofacial variation <strong>and</strong> evolution. We summarize key<br />

aspects of <strong>cranial</strong> <strong>base</strong> growth <strong>and</strong> development in <strong>primate</strong>s that are useful<br />

for formulating <strong>and</strong> testing hypotheses about the roles of the chondrocranium<br />

<strong>and</strong> basicranium in <strong>cranial</strong> growth, integration, <strong>and</strong> <strong>function</strong> in <strong>primate</strong> <strong>and</strong><br />

human evolution. We review interspecific, experimental, <strong>and</strong> ontogenetic<br />

evidence for interactions between the <strong>cranial</strong> <strong>base</strong> <strong>and</strong> brain, <strong>and</strong> between<br />

the <strong>cranial</strong> <strong>base</strong> <strong>and</strong> the face. <strong>The</strong>se interactions indicate that the <strong>cranial</strong><br />

<strong>base</strong> plays a key role in craniofacial growth, helping to integrate, spatially<br />

<strong>and</strong> <strong>function</strong>ally, different patterns of growth in various adjoining regions of<br />

the skull such as components of the brain, the eyes, the nasal cavity, the oral<br />

cavity, <strong>and</strong> the pharynx. Brain size relative to <strong>cranial</strong> <strong>base</strong> length appears to<br />

be the dominant influence on many aspects of basi<strong>cranial</strong> variation, especially<br />

the angle of the <strong>cranial</strong> <strong>base</strong> in the midsagittal plane, but other factors<br />

such as facial size, facial orientation, <strong>and</strong> posture may also be important.<br />

Major changes in <strong>cranial</strong> <strong>base</strong> shape appear to have played crucial roles in<br />

the evolution of early <strong>primate</strong>s, the origin of anthropoids, <strong>and</strong> the origin of<br />

Homo sapiens. Yrbk Phys Anthropol 43:117–169, 2000. © 2000 Wiley-Liss, Inc.<br />

TABLE OF CONTENTS<br />

Glossary, L<strong>and</strong>mark, <strong>and</strong> Measurement Definitions ....................................................... 118<br />

Introduction ......................................................................................................................... 120<br />

Anatomy <strong>and</strong> Development ................................................................................................ 120<br />

Development of the chondrocranium ............................................................................. 121<br />

Patterns <strong>and</strong> processes of basi<strong>cranial</strong> growth .............................................................. 122<br />

Antero-posterior growth .............................................................................................. 123<br />

Medio-lateral growth ................................................................................................... 125<br />

Supero-inferior growth ................................................................................................ 126<br />

Angulation .................................................................................................................... 126<br />

Associations Between Cranial Base <strong>and</strong> Brain ................................................................ 131<br />

Brain size <strong>and</strong> <strong>cranial</strong> <strong>base</strong> angle .................................................................................. 131<br />

Brain shape <strong>and</strong> <strong>cranial</strong> <strong>base</strong> angle .............................................................................. 134<br />

© 2000 WILEY-LISS, INC.


118 YEARBOOK OF PHYSICAL ANTHROPOLOGY [Vol. 43, 2000<br />

Brain volume <strong>and</strong> posterior <strong>cranial</strong> fossa shape .......................................................... 138<br />

Associations Between Cranial Base <strong>and</strong> Face .................................................................. 138<br />

Anterior <strong>cranial</strong> fossa shape <strong>and</strong> upper facial growth ................................................. 138<br />

Middle <strong>cranial</strong> fossa shape <strong>and</strong> midfacial growth ........................................................ 140<br />

Basi<strong>cranial</strong> width <strong>and</strong> overall facial shape in humans ................................................ 144<br />

<strong>The</strong> Cranial Base <strong>and</strong> Posture ........................................................................................... 148<br />

Major Unresolved Issues of Cranial Base Variation in Primate Evolution ................... 151<br />

Primate origins, encephalization, <strong>and</strong> circumorbital form .......................................... 151<br />

Anthropoid origins <strong>and</strong> <strong>cranial</strong> <strong>base</strong>-face interactions ................................................. 152<br />

Variation in hominin <strong>cranial</strong> <strong>base</strong> angle ....................................................................... 153<br />

Cranial <strong>base</strong> flexion <strong>and</strong> vocal tract shape in hominins .............................................. 154<br />

Cranial <strong>base</strong> shape <strong>and</strong> facial projection in Homo ....................................................... 156<br />

Cranial <strong>base</strong> characters in phylogenetic analyses ........................................................ 158<br />

Conclusions .......................................................................................................................... 159<br />

What major factors generate variation in the <strong>cranial</strong> <strong>base</strong>? ....................................... 160<br />

What role does the <strong>cranial</strong> <strong>base</strong> play in craniofacial integration? .............................. 160<br />

Correlation ................................................................................................................... 161<br />

Constraint .................................................................................................................... 162<br />

Ontogenetic sequence .................................................................................................. 162<br />

Future research ............................................................................................................... 163<br />

Acknowledgments ............................................................................................................... 163<br />

Literature Cited .................................................................................................................. 163<br />

GLOSSARY<br />

Basioccipital clivus: midline “plane” of the<br />

posterior <strong>cranial</strong> <strong>base</strong> formed by the superior<br />

(endo<strong>cranial</strong>) aspects of the basioccipital<br />

<strong>and</strong> the posterior sphenoid.<br />

Brain stem: the ventral parts of the brain,<br />

excluding the telencephalon. Specifically, in<br />

this paper, the brain stem consists of the<br />

medulla oblongata <strong>and</strong> mesencephalon (<br />

optic tectum <strong>and</strong> tegmentum) of Stephan et<br />

al. (1981; see also Butler <strong>and</strong> Hodos, 1996).<br />

Chondrocranium: cartilaginous precursors<br />

to the basicranium.<br />

Constraint: a limitation or bias on processes<br />

<strong>and</strong>/or patterns of evolution, growth, form,<br />

<strong>and</strong> <strong>function</strong>.<br />

Cranial <strong>base</strong> angulation: a series of events<br />

by which bone or cartilage deposition in the<br />

midline <strong>cranial</strong> <strong>base</strong> changes the angle between<br />

intersecting prechordal (see below)<br />

<strong>and</strong> postchordal (see below) lines. This causes<br />

the inferior <strong>cranial</strong> <strong>base</strong> angle to become more<br />

acute (flexion) or more obtuse (extension).<br />

Displacement: a series of events by which<br />

an osseous region “moves” relative to another<br />

osseous region through bone deposition<br />

(primary displacement), or through<br />

bone deposition in an adjoining bone (secondary<br />

displacement).<br />

Drift: a series of events by which an osseous<br />

wall “moves” relative to another anatomical<br />

region through bone deposition on one surface<br />

<strong>and</strong> bone resorption on its opposing<br />

surface.<br />

Ethmomaxillary complex: the upper part of<br />

the face, mostly comprising the ethmoid, the<br />

nasal capsule, <strong>and</strong> the maxilla.<br />

Facial projection: degree to which face<br />

projects in front of <strong>cranial</strong> <strong>base</strong>; measured<br />

here by nasion-foramen caecum.<br />

Integration: the genetic, epigenetic, or <strong>function</strong>al<br />

association among elements via “a set<br />

causal mechanisms so that change in one<br />

element is reflected by change in another”<br />

(Smith, 1996). <strong>The</strong> results of integration are<br />

most often recognized as a pattern of significant,<br />

hierarchical covariation among the<br />

components of a system.<br />

Kyphosis: angle of some aspect of facial orientation<br />

relative to the neuro- <strong>and</strong>/or basicranium,<br />

measured here using angle of facial<br />

kyphosis (AFK) for the orientation of<br />

the palate, <strong>and</strong> angle of orbit axis orientation<br />

(AOA) for orientation of the orbital axis.<br />

Planum sphenoideum: midline “plane” of<br />

the anterior <strong>cranial</strong> <strong>base</strong> from the sphenoidale<br />

(Sp) to the planum sphenoideum (PS)<br />

point (see below).


D.E. Lieberman et al.]<br />

PRIMATE CRANIAL BASE 119<br />

Postchordal <strong>cranial</strong> <strong>base</strong>: portion of the <strong>cranial</strong><br />

<strong>base</strong> posterior to the sella; frequently<br />

called the posterior <strong>cranial</strong> <strong>base</strong>.<br />

Prechordal <strong>cranial</strong> <strong>base</strong>: portion of the <strong>cranial</strong><br />

<strong>base</strong> anterior to sella; frequently called<br />

the anterior <strong>cranial</strong> <strong>base</strong>.<br />

Telencephalon: forebrain, consisting of<br />

paired olfactory lobes, the basal ganglia,<br />

<strong>and</strong> the neocortex.<br />

LANDMARK DEFINITIONS<br />

Ba, basion: midsagittal point on anterior<br />

margin of foramen magnum.<br />

CP, clival point: midline point on basioccipital<br />

clivus inferior to point at which dorsum<br />

sellae curves posteriorly.<br />

FC, foramen caecum: pit on cribriform plate<br />

between crista galli <strong>and</strong> endo<strong>cranial</strong> wall of<br />

frontal bone.<br />

H, hormion: most posterior midline point on<br />

vomer.<br />

OA: supero-inferior midpoint between superior<br />

orbital fissures <strong>and</strong> inferior rims of optic<br />

canals; for mammals without completely<br />

enclosed orbits, OA is defined as inferior rim<br />

of optic foramen.<br />

OM: supero-inferior midpoint between<br />

lower <strong>and</strong> upper orbital rims.<br />

Op, opisthion: most posterior point in foramen<br />

magnum.<br />

PMp, PM point: average of projected midline<br />

points of most anterior point on lamina<br />

of greater wings of sphenoid.<br />

PP, pituitary point: “the anterior edge of<br />

the groove for the optic chiasma, just in<br />

front of the pituitary fossa” (Zuckerman,<br />

1955).<br />

PS, planum sphenoideum point: most superior<br />

midline point on sloping surface in<br />

which cribriform plate is set.<br />

Ptm, pterygomaxillare: average of projected<br />

midline points of most inferior <strong>and</strong> posterior<br />

points on maxillary tuberosities.<br />

S, sella: center of sella turcica, independent<br />

of contours of clinoid processes.<br />

Sb, sphenobasion: midline point on sphenooccipital<br />

synchondrosis on external aspect of<br />

clivus.<br />

Sp, sphenoidale: most posterior, superior<br />

midline point of planum sphenoideum.<br />

ANGLE, LINE, AND PLANE DEFINITIONS<br />

AOA: orbital axis orientation relative to CO<br />

(Ross <strong>and</strong> Ravosa, 1993).<br />

BL1: Ba-PP PP-Sp (Ross <strong>and</strong> Ravosa,<br />

1993; Ross <strong>and</strong> Henneberg, 1995).<br />

BL2: Ba-S S-FC (Spoor, 1997).<br />

CBA1: Ba-S relative to S-FC (Lieberman<br />

<strong>and</strong> McCarthy, 1999).<br />

CBA2: Ba-S relative to Sp-PS (Lieberman<br />

<strong>and</strong> McCarthy, 1999).<br />

CBA3: Ba-CP relative to S-FC (Lieberman<br />

<strong>and</strong> McCarthy, 1999).<br />

CBA4: Ba-CP relative to Sp-PS (Lieberman<br />

<strong>and</strong> McCarthy, 1999).<br />

CO, clivus ossis occipitalis: endo<strong>cranial</strong> line<br />

from Ba to spheno-occipital synchondrosis<br />

(Ross <strong>and</strong> Ravosa, 1993).<br />

External CBA (CBA5): angle between basionsphenobasion-hormion<br />

(Lieberman <strong>and</strong> Mc-<br />

Carthy, 1999).<br />

FM, foramen magnum: Ba-Op.<br />

Forel’s axis: from most antero-inferior point<br />

on frontal lobe to most postero-inferior point<br />

on occipital lobe (Hofer, 1969).<br />

Head-neck angle: orientation of head relative<br />

to neck in locomoting animals, calculated<br />

as neck inclination orbit inclination<br />

(Strait <strong>and</strong> Ross, 1999).<br />

IRE1: cube root of endo<strong>cranial</strong> volume/BL 1<br />

(Ross <strong>and</strong> Ravosa, 1993).<br />

IRE2: cube root of neocortical volume/BL 1<br />

(Ross <strong>and</strong> Ravosa, 1993).<br />

IRE3: cube root of telencephalon volume/BL<br />

1 (Ross <strong>and</strong> Ravosa, 1993).<br />

IRE4: cube root of neocortical volume/palate<br />

length (Ross <strong>and</strong> Ravosa, 1993).<br />

IRE5: cube root of endo<strong>cranial</strong> volume/BL 2<br />

(McCarthy, 2001).<br />

Meynert’s axis: from ventral edge of junction<br />

between pons <strong>and</strong> medulla to caudal<br />

recess of interpeduncular fossa (Hofer,<br />

1969).<br />

Neck inclination: orientation of surface of<br />

neck relative to substrate (Strait <strong>and</strong> Ross,<br />

1999).<br />

NHA: neutral horizontal axis of orbits; from<br />

OM to OA (Enlow <strong>and</strong> Azuma, 1975).<br />

Orbital axis orientation: line from optic foramen<br />

through superoinferior midpoint of<br />

orbital aperture (Ravosa, 1988).<br />

Orbit inclination: orientation relative to<br />

substrate of a line joining superior <strong>and</strong> in-


120 YEARBOOK OF PHYSICAL ANTHROPOLOGY [Vol. 43, 2000<br />

ferior margins of orbits (Strait <strong>and</strong> Ross,<br />

1999).<br />

PM plane, posterior maxillary plane: from<br />

Ptm to PMp (Enlow <strong>and</strong> Azuma, 1975).<br />

INTRODUCTION<br />

<strong>The</strong> <strong>cranial</strong> <strong>base</strong> has important integrative<br />

<strong>and</strong> <strong>function</strong>al roles in the skull, many<br />

of which reflect its phylogenetic history as<br />

the oldest component of the vertebrate skull<br />

(de Beer, 1937). Architecturally, the <strong>cranial</strong><br />

<strong>base</strong> provides the platform upon which the<br />

brain grows <strong>and</strong> around which the face<br />

grows. In addition, the <strong>cranial</strong> <strong>base</strong> connects<br />

the cranium with the rest of the body:<br />

it articulates with the vertebral column <strong>and</strong><br />

the m<strong>and</strong>ible, provides conduits for all the<br />

vital neural <strong>and</strong> circulatory connections between<br />

the brain <strong>and</strong> the face <strong>and</strong> neck,<br />

houses <strong>and</strong> connects the sense organs in the<br />

skull, <strong>and</strong> forms the roof of the nasopharynx.<br />

<strong>The</strong> shape of the <strong>cranial</strong> <strong>base</strong> is therefore<br />

a multifactorial product of numerous<br />

phylogenetic, developmental, <strong>and</strong> <strong>function</strong>al<br />

interactions.<br />

<strong>The</strong> importance of the <strong>cranial</strong> <strong>base</strong> is<br />

matched by several challenges that make it<br />

difficult to study. Because the <strong>cranial</strong> <strong>base</strong><br />

is difficult to access surgically, there have<br />

been few experimental studies of <strong>cranial</strong><br />

<strong>base</strong> growth <strong>and</strong> <strong>function</strong>. Also, a large proportion<br />

of the <strong>cranial</strong> <strong>base</strong> is not only complex<br />

anatomically, but is also difficult to<br />

measure <strong>and</strong>/or see externally. In addition,<br />

the <strong>cranial</strong> <strong>base</strong> in many fossils is missing,<br />

damaged, or unobservable without special<br />

technology. However, new developmental<br />

studies, <strong>and</strong> new techniques for imaging,<br />

have led to a modest renaissance of research<br />

on <strong>cranial</strong> <strong>base</strong> morphology (reviewed in<br />

Spoor et al., 2000). In addition, new analytical<br />

techniques which quantitatively compare<br />

three-dimensional differences in form<br />

have opened up new possibilities for studying<br />

growth <strong>and</strong> variation in complex regions<br />

such as the <strong>cranial</strong> <strong>base</strong> (Cheverud <strong>and</strong><br />

Richtsmeier, 1986; Bookstein, 1991; Lele,<br />

1993; O’Higgins, 2000). Ultimately, better<br />

information about the relationships between<br />

<strong>cranial</strong> <strong>base</strong> morphology <strong>and</strong> the rest<br />

of the skull may help to resolve a number of<br />

important phylogenetic <strong>and</strong> behavioral issues<br />

throughout <strong>primate</strong> evolution.<br />

<strong>The</strong> goals of this review are to provide a<br />

background on key aspects of <strong>cranial</strong> <strong>base</strong><br />

growth <strong>and</strong> development necessary to formulate<br />

or test hypotheses about the role of<br />

the <strong>cranial</strong> <strong>base</strong> in <strong>cranial</strong> growth, integration,<br />

<strong>and</strong> <strong>function</strong>. <strong>The</strong>refore, we review recent<br />

research on <strong>cranial</strong> <strong>base</strong> variation, development,<br />

<strong>and</strong> evolution in <strong>primate</strong>s,<br />

focusing on the major dimensions of the <strong>cranial</strong><br />

<strong>base</strong> (especially width, length, <strong>and</strong> angulation<br />

in the sagittal plane). Other, more<br />

detailed aspects of <strong>cranial</strong> <strong>base</strong> anatomy<br />

<strong>and</strong> morphology, most notably the inner ear,<br />

were recently reviewed by Spoor <strong>and</strong> Zonneveld<br />

(1998) <strong>and</strong> will not be covered in this<br />

review (see also Braga et al., 1999). Where<br />

relevant, we have made an effort to include<br />

data from the few experimental studies on<br />

<strong>cranial</strong> <strong>base</strong> growth <strong>and</strong> <strong>function</strong>. Most experimental<br />

research on the mammalian<br />

skull has focused on the face <strong>and</strong> neurocranium;<br />

however, some of these studies provide<br />

indirect clues on interrelationships<br />

among the brain, <strong>cranial</strong> <strong>base</strong>, <strong>and</strong> face<br />

(e.g., Sarnat, 1988, 1999). Finally, we conclude<br />

with a short discussion of two main<br />

issues which we believe require further research<br />

to address: what factors determine<br />

most of the variation in <strong>cranial</strong> <strong>base</strong> shape<br />

among <strong>primate</strong>s, <strong>and</strong> to what extent does<br />

variation in <strong>cranial</strong> <strong>base</strong> form influence ontogenetic<br />

<strong>and</strong> interspecific patterns of variation<br />

in craniofacial morphology?<br />

ANATOMY AND DEVELOPMENT<br />

A detailed underst<strong>and</strong>ing of the series of<br />

events <strong>and</strong> underlying mechanisms that<br />

generate patterns 1 of morphological variation<br />

in the basicranium is vital for developing<br />

<strong>and</strong> testing hypotheses about the <strong>cranial</strong><br />

<strong>base</strong>’s role in craniofacial integration<br />

Please address all correspondence to: Daniel E. Lieberman,<br />

Department of Anthropology, George Washington <strong>University</strong>,<br />

2110 G St, Washington DC 20052.<br />

E-mail: danliebgwu.edu; phone: 202-994-0873; fax: 202-994-<br />

6097.<br />

1 <strong>The</strong> term “pattern” here refers to a static description of a<br />

configuration or relationship among things, whereas “process”<br />

refers to a series of events that occur during something’s formation.<br />

Note that we do not define process here as a causal mechanism.<br />

Most of the processes described here have multiple <strong>and</strong><br />

hierarchical levels of causation which merit further study.


D.E. Lieberman et al.]<br />

PRIMATE CRANIAL BASE 121<br />

Fig. 1. Chondrocranium in Homo sapiens (after<br />

Sperber, 1989). A: Superior view of chondro<strong>cranial</strong> precursors<br />

<strong>and</strong> ossification centers (after Sperber, 1989).<br />

Primordial cartilages are at right, <strong>and</strong> their <strong>cranial</strong><br />

<strong>base</strong> derivatives are on left. Note that the nasal capsule<br />

forms the ethmoid, the inferior concha, <strong>and</strong> the nasal<br />

septum; the presphenoid forms the sphenoid body; the<br />

orbitosphenoid forms the lesser wing of the sphenoid;<br />

the alisphenoid forms the greater wing of the sphenoid;<br />

the postsphenoid forms the sella turcica; the otic capsule<br />

forms the petrous temporal; the parachordal forms<br />

the basioccipital; <strong>and</strong> the occipital sclerotomes form the<br />

exoccipital. B: Lateral view of chondro<strong>cranial</strong> precursors<br />

in a fetus 8 weeks i.u.<br />

<strong>and</strong> <strong>function</strong>. So we begin with a brief summary<br />

of <strong>cranial</strong> <strong>base</strong> embryology, fetal<br />

growth, <strong>and</strong> postnatal growth. Most of the<br />

information summarized below derives from<br />

studies of human basi<strong>cranial</strong> growth <strong>and</strong><br />

development; the majority of these patterns<br />

<strong>and</strong> processes are generally applicable to all<br />

<strong>primate</strong>s, but we tried to distinguish those<br />

that are unique to humans or other species.<br />

Further information is available in Björk<br />

(1955), Ford (1958), Scott (1958), Moore <strong>and</strong><br />

Lavelle (1974), Starck (1975), Bosma (1976),<br />

Moss et al. (1982), Slavkin (1989), Sperber<br />

(1989), Enlow (1990), <strong>and</strong> Jeffery (1999), as<br />

well as the many references cited below.<br />

Development of the chondrocranium<br />

<strong>The</strong> human <strong>cranial</strong> <strong>base</strong> first appears in<br />

the second month of embryonic life as a<br />

narrow, irregularly shaped cartilaginous<br />

platform, the chondrocranium, ventral to<br />

the embryonic brain. <strong>The</strong> chondrocranium<br />

develops between the <strong>base</strong> of the embryonic<br />

brain <strong>and</strong> foregut about 28 days intra utero<br />

(i.u.) as condensations of neural crest cells<br />

(highly mobile, pluripotent neurectodermal<br />

cells that make up most of the head) <strong>and</strong><br />

paraxial mesoderm in the ectomeninx (a<br />

mesenchyme-derived membrane surrounding<br />

the brain) (Sperber, 1989). By the seventh<br />

week i.u., the ectomeninx has grown<br />

around the <strong>base</strong> of the brain <strong>and</strong> differentiated<br />

into nine groups of paired cartilagenous<br />

precursors (Fig. 1A,B) (Kjaer, 1990).<br />

From caudal to rostral these are: 1) four<br />

occipital condensations on either side of the<br />

future brain stem derived from sclerotomic<br />

portions of postotic somites; 2) a pair of<br />

parachordal cartilages on either side of the<br />

primitive notochord; 3) the otic capsules, lying<br />

lateral to the parachordal cartilages; 4)<br />

the hypophyseal (polar) cartilages which<br />

surround the anterior pituitary gl<strong>and</strong>; 5–6)<br />

the orbitosphenoids (ala orbitalis/lesser<br />

wing of sphenoid) <strong>and</strong> alisphenoids (ala<br />

temporalis/greater wings of sphenoid)<br />

which lie lateral to the hypophyseal cartilages;<br />

7–8) the trabecular cartilages which<br />

form the mesethmoid <strong>and</strong>, more laterally,<br />

the nasal capsule cartilages; <strong>and</strong> 9) the ala<br />

hypochiasmatica which, together with parts<br />

of the trabecular <strong>and</strong> orbitosphenoid cartilages,<br />

forms the presphenoid.<br />

<strong>The</strong> chondro<strong>cranial</strong> precursors anterior to<br />

the notochord (groups 5–9) derive solely<br />

from segmented neural crest tissue (somitomeres),<br />

while the posterior precursors<br />

(groups 1–4) derive from segmented mesodermal<br />

tissue (somites) (Noden, 1991; Couly<br />

et al., 1993; Le Douarin et al., 1993). Consequently,<br />

the middle of the sphenoid body<br />

(the mid-sphenoidal synchondrosis) marks<br />

the division between the anterior (prechordal)<br />

<strong>and</strong> posterior (postchordal) portions of<br />

the <strong>cranial</strong> <strong>base</strong> that are embryologically<br />

distinct. Antero-posterior specification of<br />

the segmental precursors of the <strong>cranial</strong> <strong>base</strong><br />

is complex <strong>and</strong> still incompletely known, but


122 YEARBOOK OF PHYSICAL ANTHROPOLOGY [Vol. 43, 2000<br />

appears to mostly involve the expression of<br />

the Hox <strong>and</strong> Dlx gene clusters (Lufkin et al.,<br />

1992; Robinson <strong>and</strong> Mahon, 1994; Vielle-<br />

Grosjean et al., 1997). For recent summaries<br />

of pattern formation <strong>and</strong> gene expression<br />

in the vertebrate <strong>cranial</strong> <strong>base</strong>, see<br />

Langille <strong>and</strong> Hall (1993) <strong>and</strong> Schilling <strong>and</strong><br />

Thorogood (2000).<br />

At least 41 ossification centers, which begin<br />

to appear in the chondrocranium about<br />

8 weeks i.u., are responsible for the transformation<br />

of the chondrocranium into the<br />

basicranium (Sperber, 1989; Kjaer, 1990).<br />

<strong>The</strong>se centers (Fig. 1B) form within a perforated<br />

<strong>and</strong> highly irregularly shaped platform<br />

known as the basal plate. In general,<br />

ossification begins with the mesodermally<br />

derived cartilages toward the caudal end of<br />

the chondrocranium, <strong>and</strong> proceeds rostrally<br />

<strong>and</strong> laterally, eventually forming the four<br />

major bones that comprise the <strong>primate</strong> basicranium:<br />

the ethmoid, most of the sphenoid,<br />

<strong>and</strong> parts of the occipital <strong>and</strong> temporal<br />

bones (which also include some intramembranous<br />

elements). <strong>The</strong> sequence in which<br />

the four bones of the <strong>cranial</strong> <strong>base</strong> ossify from<br />

the chondrocranium is complex, <strong>and</strong> still<br />

not entirely resolved (reviewed in Sperber,<br />

1989; Williams et al., 1995; Jeffery, 1999),<br />

but we highlight here the major steps, proceeding<br />

from caudal to rostral. <strong>The</strong> occipital<br />

comprises four bones surrounding the foramen<br />

magnum. <strong>The</strong> squamous portion is primarily<br />

intramembranous bone of the <strong>cranial</strong><br />

vault, except for the nuchal region,<br />

which ossifies endochondrally from two separate<br />

centers (Srivastava, 1992) <strong>and</strong> fuses<br />

with the lateral exoccipitals on either side of<br />

the foramen magnum that fuse with the<br />

basioccipital. <strong>The</strong> sphenoid body forms from<br />

fusion of the presphenoids <strong>and</strong> basisphenoid<br />

around the pituitary, forming the sella turcica<br />

(“Turkish saddle”). <strong>The</strong> greater <strong>and</strong><br />

lesser wings of the sphenoid develop from<br />

the fusion of the alisphenoid <strong>and</strong> orbitosphenoid<br />

cartilages to the body (Kodama,<br />

1976a–c; Sasaki <strong>and</strong> Kodama, 1976). Later,<br />

the medial <strong>and</strong> lateral pterygoid plates <strong>and</strong><br />

portions of the greater wings ossify intramembraneously.<br />

<strong>The</strong> temporals, which<br />

form much of the lateral aspect of the basicranium,<br />

develop from approximately 21 ossification<br />

centers, several of which are intramembranous,<br />

including the squamous,<br />

tympanic, <strong>and</strong> zygomatic regions (Shapiro<br />

<strong>and</strong> Robinson, 1980; Sperber, 1989). <strong>The</strong> petrous<br />

<strong>and</strong> mastoid parts of the temporal<br />

form the inner ear from the otic capsule, <strong>and</strong><br />

the styloid process of the temporal ossifies<br />

from cartilage in the second branchial arch.<br />

<strong>The</strong> ethmoid, which is entirely endochondral<br />

in origin, forms the center of the anterior<br />

<strong>cranial</strong> floor, <strong>and</strong> most of the nasal cavity<br />

from three ossification centers in the<br />

mesethmoid <strong>and</strong> nasal capsule cartilages<br />

(Hoyte, 1991). An additional cartilaginous<br />

ossification center detaches from the ectethmoid<br />

to form a separate scrolled bone, the<br />

inferior nasal concha, inside the nasal cavity.<br />

Patterns <strong>and</strong> processes of basi<strong>cranial</strong><br />

growth<br />

In order to underst<strong>and</strong> how the basicranium<br />

grows <strong>and</strong> <strong>function</strong>s during the fetal<br />

<strong>and</strong> postnatal periods, it is useful to keep in<br />

mind three important principles of basi<strong>cranial</strong><br />

development. First, the center of the<br />

basicranium (an oval-shaped region around<br />

the sphenoid body) attains adult size <strong>and</strong><br />

shape more rapidly than the anterior, posterior,<br />

<strong>and</strong> lateral portions, presumably because<br />

almost all the vital <strong>cranial</strong> nerves <strong>and</strong><br />

major vessels perforate the <strong>cranial</strong> <strong>base</strong> in<br />

this region (Figs. 1A, 2) (Sperber, 1989).<br />

Second, the prechordal (anterior) <strong>and</strong> postchordal<br />

(posterior) <strong>cranial</strong> <strong>base</strong> grow somewhat<br />

independently, perhaps reflecting<br />

their distinct embryonic origins <strong>and</strong> their<br />

different spatial <strong>and</strong> <strong>function</strong>al roles (outlined<br />

above). Third, most basi<strong>cranial</strong> growth<br />

in the three <strong>cranial</strong> fossae occurs independently<br />

(Fig. 2). <strong>The</strong> posterior <strong>cranial</strong> fossa,<br />

which houses the occipital lobes <strong>and</strong> the<br />

brain stem (the cerebellum <strong>and</strong> the medulla<br />

oblongata), is bounded laterally by the petrous<br />

<strong>and</strong> mastoid portions of the temporal<br />

bone, <strong>and</strong> anteriorly by the dorsum sellae of<br />

the sphenoid. <strong>The</strong> butterfly-shaped middle<br />

<strong>cranial</strong> fossa, which supports the temporal<br />

lobes <strong>and</strong> the pituitary gl<strong>and</strong>, is bounded<br />

posteriorly by the dorsum sellae <strong>and</strong> the<br />

petrous portions of the temporal, <strong>and</strong> anteriorly<br />

by the posterior borders of the lesser<br />

wings of the sphenoid, <strong>and</strong> by the anterior<br />

clinoid processes of the sphenoid. <strong>The</strong> ante-


D.E. Lieberman et al.]<br />

PRIMATE CRANIAL BASE 123<br />

Fig. 2. Superior view of human <strong>cranial</strong> <strong>base</strong> (after<br />

Enlow, 1990). Left: Division between anterior <strong>cranial</strong><br />

fossa (ACF), middle <strong>cranial</strong> fossa (MCF), <strong>and</strong> posterior<br />

<strong>cranial</strong> fossa (PCF). Right: Locations of major foramina<br />

(in black), <strong>and</strong> distribution of resorptive growth fields<br />

(dark, with ) <strong>and</strong> depository growth fields (light, with ).<br />

rior <strong>cranial</strong> fossa, which houses the frontal<br />

lobe <strong>and</strong> the olfactory bulbs, is bounded posteriorly<br />

by the lesser wings of the sphenoid.<br />

Following its initial formation, the <strong>cranial</strong><br />

<strong>base</strong> grows in a complex series of events,<br />

largely through displacement <strong>and</strong> drift (see<br />

Glossary). Four main types of growth occur<br />

within <strong>and</strong> between the endo<strong>cranial</strong> fossae:<br />

antero-posterior growth through displacement<br />

<strong>and</strong> drift; medio-lateral growth<br />

through displacement <strong>and</strong> drift; supero-inferior<br />

growth through drift; <strong>and</strong> angulation<br />

(primarily flexion <strong>and</strong> extension). In order<br />

to review how these types of growth occur,<br />

we will focus primarily on the sequence of<br />

events <strong>and</strong> patterns of basi<strong>cranial</strong> growth in<br />

humans <strong>and</strong> their major differences from<br />

nonhuman <strong>primate</strong>s.<br />

Antero-posterior growth. Basi<strong>cranial</strong><br />

elongation during <strong>ontogeny</strong> occurs in three<br />

ways: 1) drift at the anterior <strong>and</strong> posterior<br />

margins of the <strong>cranial</strong> <strong>base</strong>; 2) displacement<br />

in coronally oriented sutures such as the<br />

fronto-sphenoid; <strong>and</strong> 3) displacement in the<br />

midline of the <strong>cranial</strong> <strong>base</strong> from growth<br />

within the three synchondroses: the midsphenoid<br />

synchondrosis (MSS), the sphenoethmoid<br />

synchondrosis (SES), <strong>and</strong> the spheno-occipital<br />

synchondrosis (SOS). During<br />

the fetal period in both humans <strong>and</strong> nonhuman<br />

<strong>primate</strong>s, the midline anterior <strong>cranial</strong><br />

<strong>base</strong> grows in a pattern of positive allometry<br />

(mostly through ethmoidal growth) relative<br />

to the midline posterior <strong>cranial</strong> <strong>base</strong> (Ford,<br />

1956; Sirianni <strong>and</strong> Newell-Morris, 1980;<br />

Sirianni, 1985; Anagnostopolou et al., 1988;<br />

Sperber, 1989; Hoyte, 1991; Jeffrey, 1999).<br />

During fetal growth, several key differences<br />

emerge between humans <strong>and</strong> other <strong>primate</strong>s<br />

in the relative proportioning of the<br />

posterior <strong>cranial</strong> fossa (Fig. 3). In humans,<br />

antero-posterior growth in the basioccipital<br />

is proportionately less than in the exoccipital<br />

<strong>and</strong> squamous occipital posterior to the<br />

foramen magnum, whereas the pattern is<br />

apparently reversed in nonhuman <strong>primate</strong>s,<br />

with proportionately more growth in<br />

the basioccipital (Ford, 1956; Moore <strong>and</strong><br />

Lavelle, 1974). <strong>The</strong> nuchal plane rotates<br />

downward to become more horizontal in humans,<br />

but rotates in the reverse direction to<br />

become more vertical in nonhuman <strong>primate</strong>s,<br />

apparently because of a growth field<br />

reversal (Fig. 3). According to Duterloo <strong>and</strong><br />

Enlow (1970), the inside <strong>and</strong> outside of the<br />

nuchal plane in humans are resorptive <strong>and</strong><br />

depository growth fields, respectively; but in<br />

nonhuman <strong>primate</strong>s, the inside <strong>and</strong> outside<br />

of the nuchal plane are reported to be depository<br />

<strong>and</strong> resorptive growth fields, respectively.<br />

As a result, the foramen magnum<br />

lies close to the center of the<br />

basicranium in the human neonate <strong>and</strong><br />

more posteriorly in nonhuman <strong>primate</strong>s<br />

(Zuckerman, 1954, 1955; Schultz, 1955;<br />

Ford, 1956; Biegert, 1963; Crelin, 1969).<br />

Postnatally, the posterior <strong>cranial</strong> <strong>base</strong><br />

primarily elongates in the midline through<br />

deposition in the SOS <strong>and</strong> through posterior<br />

drift of the foramen magnum; more laterally,<br />

the posterior <strong>cranial</strong> fossa elongates<br />

through deposition in the occipitomastoid<br />

suture <strong>and</strong> through posterior drift. In all<br />

<strong>primate</strong>s, the basioccipital lengthens approximately<br />

twofold after birth, with rapid


124 YEARBOOK OF PHYSICAL ANTHROPOLOGY [Vol. 43, 2000<br />

Fig. 3. Midsagittal view of nonhuman <strong>primate</strong> (A) human (B), showing different patterns of drift of<br />

occipital, <strong>and</strong> position of foramen magnum, relative to overall <strong>cranial</strong> length in the Frankfurt horizontal.<br />

FM, center of the foramen magnum; OP (opisthocranium) <strong>and</strong> PR (prosthion) are, respectively the most<br />

posterior <strong>and</strong> anterior points on the skull in the Frankfurt horizontal. , depository surfaces; <br />

resorptive surfaces.<br />

growth during the neural growth period<br />

(e.g., up to approximately 6 years in humans)<br />

<strong>and</strong> some additional elongation occurring<br />

through the adolescent growth<br />

spurt (Ashton <strong>and</strong> Spence, 1958; Scott,<br />

1958; Riolo et al., 1974; Sirianni <strong>and</strong> Swindler,<br />

1979; Sirianni, 1985). <strong>The</strong> SOS contributes<br />

to roughly 70% of posterior <strong>cranial</strong><br />

<strong>base</strong> elongation in macaques (Sirianni <strong>and</strong><br />

Van Ness, 1978). <strong>The</strong> rest of posterior basi<strong>cranial</strong><br />

growth in nonhuman <strong>primate</strong>s occurs<br />

through posterior drift of the foramen<br />

magnum, which has been shown by fluorochrome<br />

dye labeling experiments to migrate<br />

caudally in nonhuman <strong>primate</strong>s through resorption<br />

at its posterior end <strong>and</strong> deposition<br />

at its anterior end (Michejda, 1971; Giles et<br />

al., 1981). In contrast, the foramen magnum<br />

remains in the center of the human skull<br />

<strong>base</strong>, roughly halfway between the most anterior<br />

<strong>and</strong> posterior points of the skull<br />

(Lugoba <strong>and</strong> Wood, 1990). <strong>The</strong> posterior <strong>cranial</strong><br />

<strong>base</strong> in H. sapiens still elongates during<br />

postnatal growth, but to a lesser degree<br />

than in nonhuman <strong>primate</strong>s.<br />

Postnatal elongation in the anterior <strong>cranial</strong><br />

<strong>base</strong> is somewhat more complex because<br />

of its multiple roles in neural <strong>and</strong><br />

facial growth. <strong>The</strong> anterior <strong>cranial</strong> <strong>base</strong><br />

(measured from sella to foramen caecum)<br />

elongates in concert with the frontal lobes of<br />

the brain, reaching approximately 95% of<br />

its adult length by the end of the neural<br />

growth period (e.g., 6 years in humans, 3<br />

years in chimpanzees, <strong>and</strong> 1.2 years in macaques)<br />

(Scott, 1958; Sirianni <strong>and</strong> Newell-<br />

Morris, 1978; Sirianni <strong>and</strong> Van Ness, 1978;<br />

Lieberman, 1998). Postnatal anterior <strong>cranial</strong><br />

<strong>base</strong> elongation can occur in the SES (in<br />

the midline), through displacement in the<br />

sphenoid-frontal suture, <strong>and</strong> through drift<br />

of the anterior margin of the frontal bone. In<br />

humans, however, the SES remains active<br />

<strong>and</strong> unfused until 6–8 years after birth,<br />

when the brain has completed most of its<br />

growth, but the SES apparently fuses near<br />

birth in nonhuman <strong>primate</strong>s (Michejda <strong>and</strong><br />

Lamey, 1971). <strong>The</strong>se differences in the timing<br />

<strong>and</strong> sequence of synchondroseal activity<br />

<strong>and</strong> fusion may be related to the different<br />

relative contributions of the lesser wings of<br />

the sphenoid <strong>and</strong> the frontal to the anterior<br />

<strong>cranial</strong> floor in humans <strong>and</strong> nonhuman <strong>primate</strong>s.<br />

Although there is some intraspecific<br />

variation, the lesser wing of the sphenoid in<br />

humans tends to comprise approximately<br />

one third of the <strong>cranial</strong> floor, extending all<br />

the way to the cribriform plate; in nonhuman<br />

<strong>primate</strong>s, the cribriform usually lies<br />

entirely within the ethmoid (Fig. 4), <strong>and</strong> the


D.E. Lieberman et al.]<br />

PRIMATE CRANIAL BASE 125<br />

Fig. 4. Superior view of <strong>cranial</strong> <strong>base</strong> in Homo sapiens (left) <strong>and</strong> Pan troglodytes (right) (after Aiello<br />

<strong>and</strong> Dean, 1991). FC, foramen, caecum; PS, planum sphenoideum point; SP, sphenoidale; S, sella. Note<br />

similar orientation of the inferior petrosal posterior surface (PPip) in the two species (data from Spoor,<br />

1997). In addition, the lesser wing of the sphenoid <strong>and</strong> the cribriform plate comprise a much greater<br />

percentage of the midline anterior <strong>cranial</strong> <strong>base</strong> in humans than in chimpanzees.<br />

lesser wing of the sphenoid makes up less<br />

than one tenth of the <strong>cranial</strong> floor (Van der<br />

Linden <strong>and</strong> Enlow, 1971; Aiello <strong>and</strong> Dean,<br />

1990; McCarthy, 2001). Differences in the<br />

sequence of synchondroseal fusion may also<br />

be related to differences in the timing <strong>and</strong><br />

nature of <strong>cranial</strong> <strong>base</strong> angulation in human<br />

vs. nonhuman <strong>primate</strong>s (Jeffery, 1999; see<br />

below).<br />

While the anterior <strong>cranial</strong> <strong>base</strong> grows<br />

solely during the neural growth phase (it<br />

reaches adult size at the same time as the<br />

brain), the more inferior portions of the<br />

anterior <strong>cranial</strong> <strong>base</strong> continue to grow as<br />

part of the face after the neural growth<br />

phase, forming the ethmomaxillary complex<br />

(Enlow, 1990). This complex grows<br />

downward <strong>and</strong> forward mostly through<br />

drift <strong>and</strong> displacement. In addition, the<br />

sphenoid sinus drifts anteriorly. Since the<br />

ethmoid (with the exception of the cribriform<br />

plate) primarily grows as part of the<br />

ethmomaxillary complex, its postnatal<br />

growth is most properly treated in a review<br />

of facial growth.<br />

Medio-lateral growth. How the <strong>cranial</strong><br />

<strong>base</strong> widens is important because of its various<br />

interactions with neuro<strong>cranial</strong> <strong>and</strong> facial<br />

shape (Lieberman et al., 2000; see below).<br />

<strong>The</strong> increases in width of the anterior<br />

<strong>and</strong> posterior <strong>cranial</strong> fossae occur primarily<br />

from drift (in which the external <strong>and</strong> internal<br />

surfaces of the squamae are depository<br />

<strong>and</strong> resorptive, respectively), <strong>and</strong> from intramembranous<br />

bone growth in sutures<br />

with some component of lateral orientation,<br />

such as the fronto-ethmoid <strong>and</strong> occipitomastoid<br />

sutures (Sperber, 1989). Lateral<br />

growth in the middle <strong>cranial</strong> fossa is slightly<br />

more complicated. <strong>The</strong> sphenoid body does<br />

not widen much (Kodama, 1976a,b; Sasaki<br />

<strong>and</strong> Kodama, 1976). Instead, most increases<br />

in middle <strong>cranial</strong> fossa width presumably<br />

occur in the spheno-temporal suture <strong>and</strong><br />

through lateral drift of the squamous portions<br />

of the sphenoid.<br />

Increases in cerebellum <strong>and</strong> brain-stem<br />

size have been implicated in changes in the<br />

orientation of the petrous pyramids (Fig. 4),<br />

which are more coronally oriented externally<br />

(but not internally) in humans than in<br />

nonhuman <strong>primate</strong>s (Dean, 1988). Spoor<br />

(1997) found that petrous pyramid orientation<br />

in a broad interspecific sample of pri-


126 YEARBOOK OF PHYSICAL ANTHROPOLOGY [Vol. 43, 2000<br />

mates was significantly negatively correlated<br />

with relative brain size (r 0.85,<br />

P 0.001) but not with the <strong>cranial</strong> <strong>base</strong><br />

angle. However, Jeffery (1999) found that<br />

petrous pyramid orientation is independent<br />

of relative brain size in fetal humans (during<br />

the second trimester).<br />

Supero-inferior growth. Most brain<br />

growth apparently causes the neurocranium<br />

<strong>and</strong> parts of the basicranium to grow<br />

superiorly, anteriorly, <strong>and</strong> laterally (de<br />

Beer, 1937). However, the endo<strong>cranial</strong> fossae<br />

also become slightly deeper through<br />

drift because most of the endo<strong>cranial</strong> floor is<br />

resorptive, while the inferior side of the basicranium<br />

is depository (Fig. 2) (Duterloo<br />

<strong>and</strong> Enlow, 1970; Enlow, 1990). <strong>The</strong> endo<strong>cranial</strong><br />

margins between the fossae that<br />

separate the different portions of the brain<br />

(the petrous portion of the temporal <strong>and</strong> the<br />

lesser wing of the sphenoid) do not drift<br />

inferiorly because they remain depository<br />

surfaces (Enlow, 1976). Differences in drift<br />

most likely reflect variation in the relative<br />

size of the components of the brain in conjunction<br />

with other spatial relationships<br />

among components of the skull. In particular,<br />

inferior drift of the anterior <strong>cranial</strong><br />

fossa is presumably minimal because it<br />

would impinge upon the orbits <strong>and</strong> nasal<br />

cavity that lie immediately below. <strong>The</strong> only<br />

exception is the cribriform plate which<br />

drifts inferiorly, slightly in humans (Moss,<br />

1963), but sometimes forming a “deep olfactory<br />

pit” in many species of nonhuman <strong>primate</strong>s<br />

(Cameron, 1930; Aiello <strong>and</strong> Dean,<br />

1990). Inferior drift of the middle <strong>cranial</strong><br />

fossa presumably reflects inferiorly directed<br />

growth of the temporal lobes, but this hypothesis<br />

has not been tested. Likewise, inferior<br />

drift in the posterior <strong>cranial</strong> fossa,<br />

which is shallow in most nonhuman <strong>primate</strong>s,<br />

is hypothesized to be a <strong>function</strong> of<br />

the size of the occipital lobes, the cerebellum,<br />

<strong>and</strong> the brain stem below the tentorium<br />

cerebelli. Note that <strong>cranial</strong> <strong>base</strong> flexion<br />

during growth, which occurs uniquely in<br />

humans (see below), complements inferior<br />

drift in the posterior <strong>cranial</strong> fossa by moving<br />

the floor of the posterior <strong>cranial</strong> fossa more<br />

below the middle <strong>cranial</strong> fossa.<br />

Angulation. Angulation of the <strong>cranial</strong><br />

<strong>base</strong> occurs when the prechordal <strong>and</strong> postchordal<br />

portions of the basicranium flex or<br />

extend relative to each other in the midsagittal<br />

plane (technically, flexion <strong>and</strong> extension<br />

describe a series of events in which the<br />

angle between the inferior or ventral surfaces<br />

of the <strong>cranial</strong> <strong>base</strong> decrease or increase,<br />

respectively). Angulation has been<br />

the subject of much research because flexion<br />

<strong>and</strong> extension of the <strong>cranial</strong> <strong>base</strong> affect the<br />

relative positions of the three endo<strong>cranial</strong><br />

fossae, thereby influencing a wide range of<br />

spatial relationships among the <strong>cranial</strong><br />

<strong>base</strong>, brain, face, <strong>and</strong> pharynx (see below).<br />

Although all measures of <strong>cranial</strong> <strong>base</strong> angle<br />

are similar in that they attempt to quantify<br />

the overall degree of angulation in the<br />

midsagittal plane between the prechordal<br />

<strong>and</strong> postchordal portions of the <strong>cranial</strong> <strong>base</strong>,<br />

there have been at least 17 different measurements<br />

used since Huxley (1867) first<br />

attempted to quantify the angle (reviewed<br />

in Lieberman <strong>and</strong> McCarthy, 1999, <strong>and</strong><br />

summarized in Table 1). Many of these angles<br />

differ considerably in how they measure<br />

the prechordal <strong>and</strong> postchordal planes<br />

<strong>and</strong>, consequently, the point of intersection<br />

between them. Figure 5 illustrates some of<br />

these angles. <strong>The</strong> postchordal plane is most<br />

commonly defined using two l<strong>and</strong>marks,<br />

usually basion <strong>and</strong> sella, or using the line<br />

created by the dorsal surface of the basioccipital<br />

clivus (the clival line). <strong>The</strong> prechordal<br />

plane has been measured in more<br />

diverse ways. Historically, the most common<br />

plane is defined by two l<strong>and</strong>marks,<br />

sella <strong>and</strong> nasion. <strong>The</strong> sella-nasion line is<br />

problematic, however, because nasion is actually<br />

part of the face <strong>and</strong> moves anteriorly<br />

<strong>and</strong> inferiorly relative to the <strong>cranial</strong> <strong>base</strong><br />

throughout the period of facial growth<br />

(Scott, 1958; Enlow, 1990). Recently, most<br />

researchers have defined the prechordal<br />

plane either from sella to the foramen caecum<br />

(a pit on the anterior end of the cribriform<br />

plate between the crista galli between<br />

frontal squama), or using the planum sphenoideum<br />

which extends from sphenoidale<br />

(the most postero-superior point on the tuberculum<br />

sellae) to the planum sphenoideum<br />

point (defined as the most anterior<br />

point on the surface of the midline anterior


D.E. Lieberman et al.]<br />

PRIMATE CRANIAL BASE 127<br />

TABLE 1. Commonly used measures of midsagittal <strong>cranial</strong> <strong>base</strong> angle<br />

Angle<br />

Posterior (P) <strong>and</strong> anterior (A)<br />

planes used<br />

References<br />

External <strong>cranial</strong> <strong>base</strong> angle P: basion-sella Björk, 1951, 1955; Stamrud, 1959<br />

Nasion-sella-basion A: sella-nasion Melsen, 1969; George, 1978<br />

L<strong>and</strong>zert’s sphenoidal angle<br />

Clivus/clival angle<br />

CBA4, planum angle<br />

P: clival plane<br />

A: ethmoidal plane (planum<br />

sphenoideum/ale)<br />

Clivus angle P: clival plane George, 1978<br />

A: sphenoidale-fronton<br />

Ethmoidal angle P: basion-sella Stamrud, 1959<br />

L<strong>and</strong>zert, 1866; Biegert, 1957;<br />

Moss, 1958; Hofer, 1957;<br />

Angst, 1967; Cartmill, 1970;<br />

Dmoch, 1975; Ross <strong>and</strong><br />

Ravosa, 1993; Ross <strong>and</strong><br />

Henneberg, 1995<br />

Internal <strong>cranial</strong> <strong>base</strong> angle<br />

A: sella-ethmoidale<br />

Spheno-ethmoidal angle P: basion-prosphenion Huxley, 1867; Topinard, 1890;<br />

Duckworth, 1904; Cameron,<br />

1924; Zuckerman, 1955<br />

Cameron’s cranio-facial axis<br />

P: basion-pituitary point<br />

A: pituitary point-nasion<br />

Cameron, 1924, 1925, 1927a,b,<br />

1930<br />

Basioccipito-septal angle P: basion-pituitary point Ford, 1956<br />

A: pituitary point-septal point<br />

Bolton’s external <strong>cranial</strong> <strong>base</strong> angle P: Bolton point-sella<br />

Brodie, 1941, 1953<br />

A: sella-nasion<br />

Anterior <strong>cranial</strong> <strong>base</strong> angle P: clival plane Scott, 1958; Cramer, 1977<br />

A: prosphenion-anterior cribriform<br />

point (ACP)<br />

Internal <strong>cranial</strong> <strong>base</strong> angle, basionsphenoidale-fronton<br />

Internal <strong>cranial</strong> <strong>base</strong> angle, basionsella-fronton<br />

Internal <strong>cranial</strong> <strong>base</strong> angle, basionsella-foramen<br />

caecum CBA1<br />

P: basion-sphenoidale<br />

A: sphenoidale-fronton<br />

P: basion-sella<br />

A: sella-fronton<br />

P: basion-sella<br />

A: sella-foramen caecum<br />

George, 1978<br />

George, 1978<br />

External <strong>cranial</strong> <strong>base</strong> angle, nasionsphenoidale-basion<br />

P: basion-sphenoidale<br />

A: sphenoidale-nasion<br />

Orbital angle P: clival plane Moss, 1958<br />

A: plane of superior orbital roof<br />

Planum angle (PANG) P: basion-sella Antón, 1989<br />

A: planum sphenoidale<br />

Orbital angle (OANG) P: basion-sella Antón, 1989<br />

A: plane of superior orbital roof<br />

Cousin et al., 1981; Spoor, 1997<br />

Lieberman <strong>and</strong> McCarthy,<br />

1999<br />

George, 1978<br />

<strong>cranial</strong> <strong>base</strong> posterior to the cribriform<br />

plate).<br />

Since different lines emphasize different<br />

aspects of <strong>cranial</strong> <strong>base</strong> anatomy, the choice<br />

of which <strong>cranial</strong> <strong>base</strong> angle to use is largely<br />

dependent on the question under study.<br />

Both postchordal lines tend to yield roughly<br />

similar results (George, 1978; Lieberman<br />

<strong>and</strong> McCarthy, 1999), but the prechordal<br />

lines can be substantially different. In particular,<br />

S-FC spans the entire length of the<br />

anterior <strong>cranial</strong> <strong>base</strong>, including the cribriform<br />

plate, whereas the planum sphenoideum<br />

does not measure the portion of the<br />

<strong>cranial</strong> <strong>base</strong> that includes the cribriform<br />

plate. Because of variation in the growth<br />

<strong>and</strong> position of the cribriform plate, these<br />

differences affect comparisons of anthropoids<br />

with strepsirrhines, or comparisons of<br />

<strong>primate</strong>s with other mammals (McCarthy,<br />

in press). In humans <strong>and</strong> some anthropoids,<br />

the cribriform plate lies in approximately<br />

the same plane as the planum sphenoideum,<br />

but in other anthropoids the cribriform<br />

plate lies in a deep olfactory pit within<br />

the ethmoidal notch of the frontal bone (Fig.<br />

6) (Aiello <strong>and</strong> Dean, 1990; Ravosa <strong>and</strong> Shea,<br />

1994). Moreover, in strepsirrhines <strong>and</strong><br />

other mammals with more divergent orbits<br />

<strong>and</strong> projecting snouts, the cribriform plate<br />

typically lies at a steep angle relative to the<br />

planum sphenoideum (Cartmill, 1970).<br />

Variations in <strong>cranial</strong> <strong>base</strong> angulation<br />

need to be considered in both comparative<br />

<strong>and</strong> ontogenetic studies. For example, it is<br />

well known that humans have a much more<br />

flexed <strong>cranial</strong> <strong>base</strong> than other <strong>primate</strong>s, but<br />

it is not well appreciated that the human


128 YEARBOOK OF PHYSICAL ANTHROPOLOGY [Vol. 43, 2000<br />

Fig. 5. Major <strong>cranial</strong> <strong>base</strong> angles used in this review, illustrated in a human cranium. CBA1 is the<br />

angle between the basion-sella (Ba-S) line <strong>and</strong> the sella-foramen (S-FC) caecum line. CBA4 is the angle<br />

between the midline of the postclival plane, <strong>and</strong> the midline of the planum sphenoideum.<br />

Fig. 6. Lateral radiographs of hemisected baboon cranium (A) <strong>and</strong> human cranium (B). Note different<br />

orientation <strong>and</strong> position of the cribriform plate (CP) relative to the planum sphenoideum (PS) in the two<br />

species.<br />

<strong>cranial</strong> <strong>base</strong> flexes postnatally, while the<br />

nonhuman <strong>primate</strong> <strong>cranial</strong> <strong>base</strong> extends<br />

postnatally, possibly at different locations<br />

(Hofer, 1960; Sirianni <strong>and</strong> Swindler, 1979;<br />

Cousin et al., 1981; Lieberman <strong>and</strong> Mc-<br />

Carthy, 1999). <strong>The</strong>se potentially nonhomologous<br />

differences exist because changes<br />

in the <strong>cranial</strong> <strong>base</strong> angle can occur through<br />

different growth processes (e.g., drift, displacement)<br />

at different locations. <strong>The</strong> actual<br />

cellular processes that result in angulation,<br />

however, are not completely understood.


D.E. Lieberman et al.]<br />

PRIMATE CRANIAL BASE 129<br />

Some researchers (Scott, 1958; Giles et al.,<br />

1981; Enlow, 1990) suggest that changes in<br />

<strong>cranial</strong> <strong>base</strong> angulation occur interstitially<br />

within synchondroses through a hinge-like<br />

action. If so, flexion would result from increased<br />

chrondrogenic activity in the superior<br />

vs. inferior aspect of the synchondrosis,<br />

while extension would result from increased<br />

chrondrogenic activity in the inferior vs. superior<br />

aspect of the synchondrosis. Experimental<br />

growth studies in macaques, which<br />

labeled growth using flurochrome dyes,<br />

show that angulation also occurs through<br />

drift in which depository <strong>and</strong> resorptive<br />

growth fields differ on either side of a synchondrosis,<br />

causing rotations around an<br />

axis through the synchondrosis (Michejda,<br />

1971, 1972a; Michejda <strong>and</strong> Lamey, 1971;<br />

Giles et al., 1981). All three synchondroses<br />

are involved in prenatal angulation (Hofer,<br />

1960; Hofer <strong>and</strong> Spatz, 1963; Sirianni <strong>and</strong><br />

Newell-Morris, 1980; Diewert, 1985; Anagastopolou<br />

et al., 1988; Sperber, 1989; van<br />

den Eynde et al., 1992); however, the extent<br />

to which each synchondrosis participates in<br />

postnatal flexion <strong>and</strong> extension is poorly<br />

known, <strong>and</strong> probably differs between humans<br />

<strong>and</strong> nonhuman <strong>primate</strong>s. <strong>The</strong> SOS,<br />

which remains active until after the eruption<br />

of the second permanent molars, is<br />

probably the most active synchondrosis in<br />

generating angulation in <strong>primate</strong>s (Björk,<br />

1955; Scott, 1958; Melsen, 1969). <strong>The</strong> MSS<br />

fuses prior to birth in humans (Ford, 1958),<br />

but may also be important in nonhuman<br />

<strong>primate</strong>s (Scott, 1958; Hofer <strong>and</strong> Spatz,<br />

1963; Michejda, 1971, 1972a; but see Lager,<br />

1958; Melsen, 1971; Giles et al., 1981). Finally,<br />

the SES fuses near birth in nonhuman<br />

<strong>primate</strong>s, <strong>and</strong> remains active only as a<br />

site of <strong>cranial</strong> <strong>base</strong> elongation in humans<br />

during the neural growth period (Scott,<br />

1958; Michejda <strong>and</strong> Lamey, 1971). Other<br />

ontogenetic changes in the <strong>cranial</strong> <strong>base</strong> angle<br />

(not necessarily involved in angulation<br />

itself) include posterior drift of the foramen<br />

magnum (see above), inferior drift of the<br />

cribriform plate relative to the anterior <strong>cranial</strong><br />

<strong>base</strong> (Moss, 1963), <strong>and</strong> remodeling of<br />

the sella turcica, which causes posterior<br />

movement of the sella (Baume, 1957; Shapiro,<br />

1960; Latham, 1972).<br />

A few experimental studies provide evidence<br />

for the presence of complex interactions<br />

between the brain <strong>and</strong> <strong>cranial</strong> <strong>base</strong><br />

synchondroses that influence variation in<br />

the <strong>cranial</strong> <strong>base</strong> angle. DuBrul <strong>and</strong> Laskin<br />

(1961), Moss (1976), Bütow (1990), <strong>and</strong> Reidenberg<br />

<strong>and</strong> Laitman (1991) all inhibited<br />

growth in the SOS in various animals<br />

(mostly rats), causing a more flexed <strong>cranial</strong><br />

<strong>base</strong>, presumably through inhibition of <strong>cranial</strong><br />

<strong>base</strong> extension. In most of these studies,<br />

experimentally induced kyphosis of the<br />

basicranium was also associated with a<br />

shorter posterior portion of the <strong>cranial</strong> <strong>base</strong>,<br />

<strong>and</strong> a more rounded neurocranium (see below).<br />

Artificial deformation of the <strong>cranial</strong><br />

vault also causes slight but significant increases<br />

in <strong>cranial</strong> <strong>base</strong> angulation (Antón,<br />

1989; Kohn et al., 1993). However, no controlled<br />

experimental studies have yet examined<br />

disruptions to the other <strong>cranial</strong> <strong>base</strong><br />

synchondroses. In addition, there have been<br />

few controlled studies of the effect of increasing<br />

brain size on <strong>cranial</strong> <strong>base</strong> angulation.<br />

In one classic experiment, Young<br />

(1959) added sclerosing fluid into the <strong>cranial</strong><br />

cavity in growing rats, which caused<br />

enlargement of the neurocranium with little<br />

effect on angular relationships in the <strong>cranial</strong><br />

<strong>base</strong>. Additional evidence for some degree<br />

of independence between the brain <strong>and</strong><br />

<strong>cranial</strong> <strong>base</strong> during development is provided<br />

by microcephaly <strong>and</strong> hydrocephaly, in<br />

which <strong>cranial</strong> <strong>base</strong> angles tend to be close to<br />

those of humans with normal encephalization<br />

(Moore <strong>and</strong> Lavelle, 1974; Sperber,<br />

1989).<br />

Important differences in <strong>cranial</strong> <strong>base</strong> angulation<br />

among <strong>primate</strong>s exist in terms of<br />

the ontogenetic pattern of flexion <strong>and</strong>/or extension,<br />

which presumably result from differences<br />

in the rate, timing, duration, <strong>and</strong><br />

sequence of the growth processes outlined<br />

above. Jeffery (1999) suggested that, prenatally,<br />

the basicranium in humans initially<br />

flexes rapidly during the period of rapid<br />

hindbrain growth in the first trimester, remains<br />

fairly stable during the second trimester,<br />

<strong>and</strong> then extends during the third<br />

trimester in conjunction with facial extension,<br />

even while the brain is rapidly increasing<br />

in size relative to the rest of the cranium<br />

(see also Björk, 1955; Ford, 1956; Sperber,


130 YEARBOOK OF PHYSICAL ANTHROPOLOGY [Vol. 43, 2000<br />

Fig. 7. Changes in the angle of the <strong>cranial</strong> <strong>base</strong> (CBA1 <strong>and</strong> CBA4) in Homo sapiens (left) <strong>and</strong> Pan<br />

troglodytes (right) (after Fig. 6 in Lieberman <strong>and</strong> McCarthy, 1999). For both species, means (circles) <strong>and</strong><br />

st<strong>and</strong>ard deviations (whiskers) are summarized by the mean chronological age of each dental stage. Note<br />

that the human <strong>cranial</strong> <strong>base</strong> flexes rapidly during stage I <strong>and</strong> then remains stable, whereas the<br />

chimpanzee <strong>cranial</strong> <strong>base</strong> extends gradually through stage V. *P 0.05.<br />

1989; van den Eynde et al., 1992). Little is<br />

known about prenatal <strong>cranial</strong> <strong>base</strong> angulation<br />

in nonhuman <strong>primate</strong>s; however, major<br />

differences between humans <strong>and</strong> nonhumans<br />

appear during the fetal <strong>and</strong> postnatal<br />

periods of growth. As Figure 7 illustrates,<br />

the human <strong>cranial</strong> <strong>base</strong> flexes rapidly after<br />

birth, almost entirely prior to 2 years of age,<br />

<strong>and</strong> well before the brain has ceased to exp<strong>and</strong><br />

appreciably (Lieberman <strong>and</strong> Mc-<br />

Carthy, 1999). In contrast, the nonhuman<br />

<strong>primate</strong> <strong>cranial</strong> <strong>base</strong> extends gradually after<br />

birth throughout the neural <strong>and</strong> facial<br />

growth periods, culminating in an accelerated<br />

phase during the adolescent growth<br />

spurt (Hofer, 1960; Heintz, 1966; Sirianni<br />

<strong>and</strong> Swindler, 1979; Cousin et al., 1981;<br />

Lieberman <strong>and</strong> McCarthy, 1999).<br />

<strong>The</strong> polyphasic <strong>and</strong> multifactorial nature<br />

of <strong>cranial</strong> <strong>base</strong> angulation during <strong>ontogeny</strong><br />

<strong>and</strong> the ontogenetic contrasts between the<br />

relative contribution of underlying factors<br />

in humans <strong>and</strong> nonhumans provide some<br />

clues for underst<strong>and</strong>ing the complex, multiple<br />

interactions between <strong>cranial</strong> <strong>base</strong> angulation,<br />

encephalization, <strong>and</strong> facial growth.<br />

<strong>The</strong>re is abundant evidence <strong>base</strong>d on interspecific<br />

studies (see below) that variation in<br />

<strong>cranial</strong> <strong>base</strong> angle in <strong>primate</strong>s is associated<br />

in part with variation in brain volume relative<br />

to the length of the <strong>cranial</strong> <strong>base</strong>. However,<br />

relative encephalization accounts for<br />

only 36% of the variation in <strong>cranial</strong> <strong>base</strong><br />

angle among anthropoids, <strong>and</strong> both interspecific<br />

<strong>and</strong> ontogenetic data suggest that a<br />

large proportion of the variation in <strong>cranial</strong>


D.E. Lieberman et al.]<br />

PRIMATE CRANIAL BASE 131<br />

<strong>base</strong> angle among <strong>primate</strong>s must also be<br />

related to variation in facial growth, orbit<br />

orientation, <strong>and</strong> relative orbit size (Ross <strong>and</strong><br />

Ravosa, 1993; Ravosa et al., 2000a). As<br />

noted above, the ontogenetic pattern of prenatal<br />

<strong>cranial</strong> <strong>base</strong> angulation in humans is<br />

largely unrelated to the rate at which the<br />

brain exp<strong>and</strong>s (Jeffery, 1999). In addition,<br />

the nonhuman <strong>primate</strong> <strong>cranial</strong> <strong>base</strong> angle<br />

(regardless of whether the cribriform plate<br />

is included in the measurement) mostly extends<br />

during the period of facial growth,<br />

after the brain has ceased to exp<strong>and</strong><br />

(Lieberman <strong>and</strong> McCarthy, 1999). <strong>The</strong>refore,<br />

we will next explore in greater depth<br />

the relationship between <strong>cranial</strong> <strong>base</strong> angle,<br />

brain size, relative orbit size <strong>and</strong> position,<br />

facial orientation, <strong>and</strong> other factors such as<br />

pharyngeal shape <strong>and</strong> facial projection.<br />

ASSOCIATIONS BETWEEN CRANIAL<br />

BASE AND BRAIN<br />

Because of the close relationship between<br />

the brain <strong>and</strong> the <strong>cranial</strong> <strong>base</strong> during development<br />

(see above), the hypothesis that<br />

brain size <strong>and</strong> shape influence basi<strong>cranial</strong><br />

morphology is an old <strong>and</strong> persistent one.<br />

<strong>The</strong> bones of the <strong>cranial</strong> cavity, including<br />

the <strong>cranial</strong> <strong>base</strong>, are generally known to<br />

conform to the shape of the brain, but the<br />

specifics of this relationship <strong>and</strong> any reciprocal<br />

effects of <strong>cranial</strong> <strong>base</strong> size <strong>and</strong> shape<br />

on brain morphology remain unclear. For<br />

example, the human basicranium is flexed<br />

when it first appears in weeks 5 <strong>and</strong> 6 because<br />

in the fourth week, the neural tube<br />

bends ventrally at the cephalic flexure<br />

(O’Rahilly <strong>and</strong> Müller, 1994). <strong>The</strong> parachordal<br />

condensations caudal to the cephalic<br />

flexure are therefore in a different<br />

anatomical plane than the more rostral prechordal<br />

condensations (which develop by<br />

week 7). However, as noted above, it is difficult<br />

to attribute many of the subsequent<br />

changes in prenatal chondro<strong>cranial</strong> or basi<strong>cranial</strong><br />

angulation (or other measures of the<br />

<strong>base</strong>) as responses solely to changes in brain<br />

morphology.<br />

Here we review several key aspects of the<br />

association between brain <strong>and</strong> <strong>cranial</strong> <strong>base</strong><br />

morphology, as derived from interspecific<br />

analyses of adult specimens. Structural relationships<br />

between the <strong>cranial</strong> <strong>base</strong> <strong>and</strong><br />

the face are discussed below.<br />

Brain size <strong>and</strong> <strong>cranial</strong> <strong>base</strong> angle<br />

Numerous anatomists have posited a relationship<br />

between brain size <strong>and</strong> basi<strong>cranial</strong><br />

angle (e.g., Virchow, 1857; Ranke, 1892;<br />

Cameron, 1924; Bolk, 1926; Dabelow, 1929,<br />

1931; Biegert, 1957, 1963; Delattre <strong>and</strong> Fenart,<br />

1963; Hofer, 1969; Gould, 1977; Ross<br />

<strong>and</strong> Ravosa, 1993; Ross <strong>and</strong> Henneberg,<br />

1995; Spoor, 1997; Strait, 1999; Strait <strong>and</strong><br />

Ross, 1999; McCarthy, 2001). <strong>The</strong> most<br />

widely accepted of these hypotheses is that<br />

the angle of the midline <strong>cranial</strong> <strong>base</strong> in the<br />

sagittal plane correlates with the volume of<br />

the brain relative to basi<strong>cranial</strong> length<br />

(DuBrul <strong>and</strong> Laskin, 1961; Vogel, 1964;<br />

Riesenfeld, 1969; Gould, 1977). This hypothesis<br />

is supported by independent analyses of<br />

different measures of basi<strong>cranial</strong> flexion<br />

across several interspecific samples of <strong>primate</strong>s<br />

(Ross <strong>and</strong> Ravosa, 1993; Spoor, 1997;<br />

McCarthy, 2001) (Fig. 8): the adult midline<br />

<strong>cranial</strong> <strong>base</strong> is significantly <strong>and</strong> predictably<br />

more flexed in species with larger endo<strong>cranial</strong><br />

volumes relative to basi<strong>cranial</strong> length.<br />

In particular, the analysis by Ross <strong>and</strong> Ravosa<br />

(1993) of a broad interspecific sample<br />

of <strong>primate</strong>s found that the correlation coefficient<br />

between relative encephalization<br />

(IRE1, see below) <strong>and</strong> <strong>cranial</strong> <strong>base</strong> angle<br />

(CBA4, see below) was 0.645 (P 0.001),<br />

explaining approximately 40% of the variation<br />

in <strong>cranial</strong> <strong>base</strong> angle.<br />

Attempts to extend this relationship to<br />

hominins have proved controversial. Ross<br />

<strong>and</strong> Henneberg (1995) reported that Homo<br />

sapiens have less flexed basicrania than<br />

predicted by either haplorhine or <strong>primate</strong><br />

regressions. <strong>The</strong>y posited that spatial constraints<br />

limit the degree of flexion possible,<br />

<strong>and</strong> that humans accommodate further<br />

brain expansion relative to <strong>cranial</strong> <strong>base</strong><br />

length through means other than flexion,<br />

such as superior, posterior, <strong>and</strong> lateral neuro<strong>cranial</strong><br />

expansion. In contrast, Spoor<br />

(1997), using different measures of flexion<br />

<strong>and</strong> relative brain size taken on a different<br />

sample, found H. sapiens to have the degree<br />

of flexion expected for its relative brain size.<br />

Spoor (1997) used the angle basion-sellaforamen<br />

caecum (CBA1) to quantify basi-


132 YEARBOOK OF PHYSICAL ANTHROPOLOGY [Vol. 43, 2000<br />

Fig. 8. Bivariate plot of CBA4 against IRE5. <strong>The</strong>se variables are significantly correlated across<br />

Primates (r 0.621; P 0.05) <strong>and</strong> Haplorhini (r 0.636; P 0.05).<br />

<strong>cranial</strong> flexion, <strong>and</strong> the length of these line<br />

segments (thereby including cribriform<br />

plate length) to quantify basi<strong>cranial</strong> length<br />

(BL2), whereas Ross <strong>and</strong> Henneberg (1995)<br />

measured flexion using CBA4 <strong>and</strong> relative<br />

brain size using IRE1 (endo<strong>cranial</strong> volume<br />

0.33 /BL1).<br />

<strong>The</strong> two most likely sources of the discrepancy<br />

between the results of Spoor (1997) vs.<br />

Ross <strong>and</strong> Henneberg (1995) were the different<br />

measures <strong>and</strong> different samples. Mc-<br />

Carthy (2001) has since investigated the influence<br />

of different measures, noting that<br />

the measure of basi<strong>cranial</strong> length by Ross<br />

<strong>and</strong> Henneberg (1995) excluded the horizontally<br />

oriented cribriform plate that contributes<br />

to basi<strong>cranial</strong> length in anthropoids<br />

more than in strepsirrhines. McCarthy<br />

(2001) also demonstrated that the frontal<br />

bone contributes less to midline <strong>cranial</strong> <strong>base</strong><br />

length in hominoids, especially humans,<br />

causing BL1 to underestimate midline basi<strong>cranial</strong><br />

length relative to endo<strong>cranial</strong> volume<br />

compared to other anthropoids. However,<br />

the data sets of both McCarthy (2000)<br />

<strong>and</strong> Spoor (1997) were small (n 17 species)<br />

in comparison with that of Ross <strong>and</strong><br />

Henneberg (1995) (n 64 species). We<br />

therefore reanalyzed the relationships between<br />

flexion <strong>and</strong> relative brain size in a<br />

large interspecific <strong>primate</strong> sample, utilizing<br />

both CBA1 <strong>and</strong> CBA4 as measures of flexion<br />

<strong>and</strong> IRE5 as a measure of relative brain<br />

size. 2 IRE5 incorporates the more appropriate<br />

basi<strong>cranial</strong> length that includes cribriform<br />

plate length (see Glossary <strong>and</strong> Measurement<br />

Definitions). <strong>The</strong> human value for<br />

CBA1 falls within the 95% confidence limits<br />

of the value predicted for an anthropoid of<br />

its relative brain size, but the human value<br />

for CBA4 does not. <strong>The</strong>se results corroborate<br />

those of McCarthy (2001): the degree of<br />

basi<strong>cranial</strong> flexion in humans is not significantly<br />

less than expected using CBA1, but<br />

is less than expected using CBA4. Thus humans<br />

may or may not have the degree of<br />

2 Measurements were taken on radiographs of nonhuman <strong>primate</strong>s<br />

from Ravosa (1991b) <strong>and</strong> Ross <strong>and</strong> Ravosa (1993), <strong>and</strong> on<br />

98 Homo sapiens from Ross <strong>and</strong> Henneberg (1995). RMA slopes<br />

were calculated for nonhominin <strong>primate</strong>s, <strong>and</strong> the 95% bootstrap<br />

confidence limits for the value of y predicted for humans were<br />

calculated according to Joliceur <strong>and</strong> Mosiman (1968), using software<br />

written by Tim Cole.


D.E. Lieberman et al.]<br />

PRIMATE CRANIAL BASE 133<br />

TABLE 2. Variance components for indices of relative brain size <strong>and</strong> measures of flexion across <strong>primate</strong>s 1<br />

Level<br />

N<br />

IRE1 IRE5 CBA4 CBA1<br />

% N eff % N eff % N eff % N eff<br />

Infraorder 2 12 0.24 48 0.94 0 0 52 1.05<br />

Superfamily 6 22 1.34 10 0.62 34 2.07 0 0.00<br />

Family 15 54 8.16 16 2.44 44 6.56 35 5.19<br />

Genus 60 13 7.81 22 12.99 17 10.26 8 4.74<br />

Species 62 1 0.78 3 2.38 5 3.10 5 3.25<br />

Total N eff 18.33 17.38 21.99 14.23<br />

df 16.00 15.00 20.00 12.00<br />

1 Maximum likelihood variance components were calculated using mainframe SAS (proc varcomp). %, percentage of variance at<br />

each taxonon 1 mic level; N eff , effective N at each level; Total N eff , total effective N for each variable. Bivariate comparisons utilize<br />

lowest N eff of the pair.<br />

TABLE 3. Correlation coefficients for <strong>primate</strong>s <strong>and</strong> haplorhini 1<br />

Variables CBA4 N (P) df eff (P) CBA1 N (P) df eff (P)<br />

Primates<br />

IRE1 0.783 60 (0.001) 14 (0.01) 0.672 60 (0.001) 10 (0.05)<br />

IRE5 0.621 60 (0.001) 13 (0.05) 0.790 62 (0.001) 10 (0.01)<br />

Haplorhini<br />

IRE1 0.813 51 (0.001) 9 (0.01) 0.641 51 (0.001) 9 (0.05)<br />

IRE5 0.636 51 (0.001) 10 (0.05) 0.548 51 (0.001) 14 (0.05)<br />

1 N, total N; d eff ,N eff 2, effective degrees of freedom for each comparison. Bivariate comparisons utilize lowest N eff of the pair.<br />

flexion expected for their relative brain size,<br />

depending on which measures are used.<br />

One problem with the above studies is<br />

that they do not consider the potential role<br />

of phylogenetic effects on these correlations<br />

(Cheverud et al., 1985; Felsenstein, 1985).<br />

Accordingly, the above data were reanalyzed<br />

using the method of Smith (1994) for<br />

adjusting degrees of freedom. 3 Table 2 presents<br />

the percentage of total variance distributed<br />

at each taxonomic level within the<br />

order Primates. Table 3 presents the correlation<br />

coefficients for comparisons of CBA<br />

<strong>and</strong> IRE for Primates <strong>and</strong> Haplorhini, along<br />

with sample sizes, degrees of freedom adjusted<br />

for phylogeny (df eff ), <strong>and</strong> their associated<br />

P-values (for which strepsirrhines<br />

had no significant correlations). Examination<br />

of the variance components in Table 2<br />

shows that the relationship between <strong>cranial</strong><br />

<strong>base</strong> angle <strong>and</strong> relative brain size is subject<br />

to significant phylogenetic effects. However,<br />

3 Unlike methods such as “independent contrasts” (e.g., Purvis<br />

<strong>and</strong> Rambaut, 1995), Smith’s method uses values for variables in<br />

just terminal taxa, therefore avoiding potentially spurious estimates<br />

of these values for ancestral nodes. Smith’s method is also<br />

more robust when five or fewer taxonomic levels are considered<br />

<strong>and</strong> is less affected by arbitrary taxonomic groupings (Nunn,<br />

1995). We also used the maximum likelihood method for calculating<br />

variance components rather than a nested ANOVA<br />

method because maximum likelihood does not generate negative<br />

variance components.<br />

the method of Smith (1994) is conservative<br />

<strong>and</strong> the correlations that survive these corrected<br />

degrees of freedom are robust, although<br />

of relatively low magnitude. Across<br />

<strong>primate</strong>s <strong>and</strong> haplorhines, both CBA4 <strong>and</strong><br />

CBA1 are significantly correlated (P 0.05)<br />

with IRE5 (Fig. 8; Table 2). <strong>The</strong>se results<br />

corroborate the results of Ross <strong>and</strong> Ravosa<br />

(1993), but with a more appropriate measure<br />

of basi<strong>cranial</strong> length incorporated into<br />

the measure of relative brain size. This confirms<br />

that brain size relative to basi<strong>cranial</strong><br />

length is significantly correlated with basi<strong>cranial</strong><br />

flexion, but that the correlations are<br />

not particularly strong, indicating that<br />

there may be other important influences on<br />

the degree of basi<strong>cranial</strong> flexion (see below).<br />

Most hypotheses explaining basi<strong>cranial</strong><br />

flexion have focused on increases in relative<br />

brain size as the variable driving flexion.<br />

However, as Strait (1999) has noted, it is<br />

important to consider the scaling relationships<br />

of basi<strong>cranial</strong> length <strong>and</strong> brain size.<br />

Using interspecific data from Ross <strong>and</strong> Ravosa<br />

(1993) in conjunction with other studies,<br />

Strait (1999) found that basi<strong>cranial</strong><br />

length scales with negative allometry<br />

against body mass <strong>and</strong> telencephalon volume<br />

(results confirmed here using BL2 instead<br />

of BL1; see Table 4), <strong>and</strong> BL2 also


134 YEARBOOK OF PHYSICAL ANTHROPOLOGY [Vol. 43, 2000<br />

TABLE 4. Reduced major axis regression equations for scaling of basi<strong>cranial</strong> dimensions against<br />

neural variables 1<br />

Comparison Intercept Slope<br />

95% CI<br />

of slope Scaling 3 N r<br />

BL2 vs.<br />

Body mass 0.908 0.654 0.035 ve 61 0.978<br />

Telencephalon 0.541 0.742 0.091 ve 33 0.960<br />

Brain stem 0.404 1.123 0.100 ve/Iso 33 0.975<br />

Neuro<strong>cranial</strong> volume 1.270 0.734 0.056 ve 62 0.962<br />

Palate (Pros-PNS) 1.394 0.642 0.032 ve 58 0.929<br />

Anterior face (Pros-Nas) 1.623 0.577 0.028 ve 58 0.933<br />

Upper toothroow 1.425 0.666 0.035 ve 58 0.921<br />

Geometric mean 2 0.951 0.750 0.023 ve 58 0.972<br />

Ba-S vs.<br />

Body mass 0.380 0.759 0.056 ve 63 0.958<br />

Brain stem 0.226 1.354 0.161 ve 33 0.946<br />

Telencephalon 0.313 0.908 0.135 Iso 33 0.924<br />

Cerebellum 0.105 0.953 0.111 Iso 33 0.948<br />

Infratentorial brain 0.064 1.044 0.123 Iso 33 0.947<br />

Palate (Pros-PNS) 0.595 0.757 0.040 ve 58 0.917<br />

Anterior face (Pros-Nas) 0.862 0.681 0.037 ve 58 0.914<br />

Upper toothroow 0.626 0.787 0.043 ve 58 0.924<br />

Geometric mean 2 0.067 0.886 0.034 ve 58 0.957<br />

S-FC vs.<br />

Ba-S 0.473 0.759 0.063 ve 33 0.949<br />

Body mass 0.748 0.587 0.040 ve 56 0.965<br />

Brain stem 0.249 1.070 0.101 Iso 33 0.971<br />

Telencephalon 0.378 0.715 0.082 ve 33 0.963<br />

Cerebellum 0.516 0.749 0.079 ve 33 0.964<br />

Infratentorial brain 0.381 0.822 0.085 ve 33 0.965<br />

Palate (Pros-PNS) 1.382 0.581 0.032 ve 58 0.912<br />

Anterior face (Pros-Nas) 1.586 0.523 0.027 ve 58 0.922<br />

Upper toothroow 1.406 0.604 0.035 ve 58 0.900<br />

Geometric mean 2 0.977 0.680 0.026 ve 58 0.958<br />

1 All volumetric <strong>and</strong> mass variables converted to cube roots. All calculations in log-log space (<strong>base</strong> 10). Calculations with facial<br />

dimensions do not include humans; all other calculations include humans.<br />

2 Geometric mean of the following measures: palate length, palate breadth, anterior face length, maxillary postcanine toothrow<br />

length, outer biorbital breadth, bizygomatic breadth, basi<strong>cranial</strong> length, <strong>and</strong> lower skull length. Measures of brain volume from<br />

Stephan et al. (1981), measures of body mass from Smith <strong>and</strong> Jungers (1994), <strong>and</strong> measures of facial size derive from the same skulls<br />

from which the radiographs were taken.<br />

3 ve, negative; ve, positive; Iso, isometric.<br />

scales with negative allometry against measures<br />

of facial size (Table 4). <strong>The</strong> only variable<br />

that scales close to isometry with basi<strong>cranial</strong><br />

length is brain-stem volume, which<br />

is isometric with BL1 <strong>and</strong> scales close to<br />

isometry with BL2 (Table 4). Intuitively this<br />

isometry makes sense, because the brain<br />

stem rests on the basicranium. However,<br />

the posterior basicranium (Ba-S), which<br />

predominantly underlies the brain stem,<br />

scales with positive allometry to brain-stem<br />

volume, whereas the anterior <strong>cranial</strong> <strong>base</strong><br />

length scales isometrically with brain-stem<br />

volume. Notably, the anterior <strong>cranial</strong> <strong>base</strong><br />

always shows lower slopes than the posterior<br />

<strong>cranial</strong> <strong>base</strong>, suggesting that the strong<br />

negative allometry of <strong>cranial</strong> <strong>base</strong> length<br />

relative to neural variables, <strong>and</strong> possibly<br />

flexion, is disproportionately attributable to<br />

relative shortening of the anterior <strong>cranial</strong><br />

<strong>base</strong>.<br />

Brain shape <strong>and</strong> <strong>cranial</strong> <strong>base</strong> angle<br />

Given the close anatomical relationship<br />

between the brain <strong>and</strong> basicranium, it<br />

seems intuitive that the shapes of the two<br />

should be related, but testing this hypothesis<br />

in a controlled fashion has been difficult.<br />

<strong>The</strong> best evidence for the presence of interactions<br />

between brain shape <strong>and</strong> the <strong>cranial</strong><br />

<strong>base</strong> came from studies of artificial <strong>cranial</strong><br />

vault deformation <strong>and</strong> from the effects of<br />

closing various <strong>cranial</strong> vault sutures on the<br />

<strong>cranial</strong> <strong>base</strong>. <strong>The</strong> effects of head-binding are<br />

difficult to interpret without precise data on<br />

the timing <strong>and</strong> forces used to deform the<br />

skull. Nevertheless, antero-posterior headbinding<br />

tends to cause lateral expansion of


D.E. Lieberman et al.]<br />

PRIMATE CRANIAL BASE 135<br />

Fig. 9. Bivariate plot of the measure by Hofer (1969)<br />

of basi<strong>cranial</strong> flexion <strong>and</strong> CBA4 against his measure of<br />

brain flexion. Hofer’s brain angle is the anterior angle<br />

between Forel’s axis, from the most antero-inferior<br />

point on the frontal lobe to the most postero-inferior<br />

point on the occipital lobe, <strong>and</strong> Meynert’s axis, from the<br />

ventral edge of the junction between the pons <strong>and</strong> medulla<br />

to the caudal recess of the interpeduncular fossa.<br />

<strong>The</strong> data of Hofer (1969) measure anterior angles between<br />

lines, rather than the inferior angles favored by<br />

recent workers (e.g., Ross <strong>and</strong> Ravosa, 1993). <strong>The</strong> plotted<br />

data therefore represent the complement of Hofer’s<br />

angles.<br />

the <strong>cranial</strong> <strong>base</strong> along with a slight increase<br />

in CBA; conversely, annular head-binding<br />

tends to cause medio-lateral narrowing <strong>and</strong><br />

antero-posterior elongation of the <strong>cranial</strong><br />

<strong>base</strong>, also with a slight increase in CBA<br />

(Antón, 1989; Cheverud et al., 1992; Kohn et<br />

al., 1993). Natural or experimentally induced<br />

premature closure of sutures (synostoses)<br />

in the <strong>cranial</strong> vault have similarly<br />

predictable effects. For example, bilateral<br />

coronal synostoses cause antero-posterior<br />

shortening of the <strong>cranial</strong> <strong>base</strong> (Babler, 1989;<br />

David et al., 1989), <strong>and</strong> unilateral coronal<br />

synostoses (plagiocephaly) cause marked<br />

asymmetry in the <strong>cranial</strong> vault, <strong>cranial</strong><br />

<strong>base</strong>, <strong>and</strong> face.<br />

Interspecific analyses of the relationship<br />

between brain shape <strong>and</strong> <strong>cranial</strong> <strong>base</strong> shape<br />

in <strong>primate</strong>s are rare. Hofer (1965, 1969)<br />

measured the orientation of the cerebral<br />

hemispheres relative to the brain stem in<br />

<strong>primate</strong>s using two axes: Forel’s axis, from<br />

the most antero-inferior point on the frontal<br />

lobe to the most postero-inferior point on the<br />

occipital lobe, measuring the orientation of<br />

the inferior surface of the cerebral hemispheres;<br />

<strong>and</strong> Meynert’s axis, from the ventral<br />

edge of the junction between the pons<br />

<strong>and</strong> medulla to the caudal recess of the interpeduncular<br />

fossa, quantifying the orientation<br />

of the brain stem. Hofer (1965) also<br />

measured the angle of the midline <strong>cranial</strong><br />

<strong>base</strong> using a modified version of the angle of<br />

L<strong>and</strong>zert (1866), similar to the CBA4 used<br />

by Ross <strong>and</strong> Ravosa (1993).<br />

Figure 9, a plot of the measure by Hofer of<br />

basi<strong>cranial</strong> angle against his measure of<br />

brain angle, illustrates that these variables<br />

are highly correlated <strong>and</strong> scale isometrically<br />

with each other (i.e., have a slope of 1.0). As<br />

the cerebrum flexes on the brain stem, the<br />

planum sphenoideum flexes relative to the<br />

clivus. <strong>The</strong> explanation of Hofer (1969) for<br />

this phenomenon is that the telencephalon<br />

becomes more spherical as it enlarges, to<br />

minimize surface area relative to volume.<br />

An alternative hypothesis is that increasing<br />

the antero-posterior diameter of the head<br />

“would be disastrous, making larger animals<br />

unusually long-headed, <strong>and</strong> would pro-


136 YEARBOOK OF PHYSICAL ANTHROPOLOGY [Vol. 43, 2000<br />

TABLE 5. Correlation coefficients for flexion vs. measures of relative brain size<br />

N N eff (df) CBA4 CBA1<br />

Neocortex volume/BL2<br />

Primates 29 12 (10) 0.613 (0.05) 0.759 (0.01)<br />

Haplorhines 22 9 (7) 0.636 (ns) 0.660 (ns)<br />

Strepsirhines 7 6 (4) 0.282 (ns) 0.771 (ns)<br />

Telencephalon/brain-stem volume<br />

Primates 29 9 (7) 0.744 (0.05) 0.663 (ns)<br />

Haplorhines 19 6 (4) 0.726 (ns) 0.643 (ns)<br />

Strepsirhines 9 7 (5) 0.760 (0.05) 0.286 (ns)<br />

duce serious problems for balancing the<br />

skull on the skeleton” (Jerison, 1982, p. 82).<br />

In the context of such spatial constraints<br />

(i.e., limited cerebral diameter <strong>and</strong> tendency<br />

to sphericity), increased cerebrum size can<br />

only be accommodated by exp<strong>and</strong>ing the<br />

<strong>cranial</strong> <strong>base</strong> inferiorly, posteriorly, or anteriorly,<br />

thereby necessarily flexing the brain<br />

<strong>and</strong> the basicranium. <strong>The</strong> hypothesis of<br />

Jerison (1982) is refuted by animals such as<br />

camels, llamas, <strong>and</strong> giraffes, which have<br />

long heads on relatively orthograde necks;<br />

<strong>and</strong> as discussed below, there is no convincing<br />

evidence that head <strong>and</strong> neck posture are<br />

significant influences on basi<strong>cranial</strong> angle<br />

in <strong>primate</strong>s (Strait <strong>and</strong> Ross, 1999).<br />

Biegert (1963) made a claim similar to<br />

that of Hofer (1969), in arguing that increases<br />

in <strong>primate</strong> brain size relative to <strong>cranial</strong><br />

<strong>base</strong> size, as well as increases in “neopallium”<br />

(i.e., neocortex) size relative to<br />

other parts of the brain, produced a rounder<br />

brain such that “adaptations in the structure<br />

of the cranium accompanied these<br />

changes in the size <strong>and</strong> shape of the brain”<br />

(Biegert, 1963, p. 120). <strong>The</strong> predicted<br />

changes in skull shape include increased<br />

vaulting of the frontal <strong>and</strong> occipital bones<br />

<strong>and</strong> increased basi<strong>cranial</strong> flexion. Ross <strong>and</strong><br />

Ravosa (1993) evaluated the hypothesis of<br />

Biegert (1963) by calculating correlation coefficients<br />

between <strong>cranial</strong> <strong>base</strong> angle<br />

(CBA4) <strong>and</strong> the ratio of neocortical volume<br />

0.33 /basi<strong>cranial</strong> length. Although they<br />

found a significant relationship across <strong>primate</strong>s<br />

<strong>and</strong> haplorhines, the correlation coefficients<br />

were low (around 0.5). Recalculation<br />

of these correlation coefficients using BL2<br />

as a measure of basi<strong>cranial</strong> length produces<br />

significant correlations across <strong>primate</strong>s,<br />

haplorhines, <strong>and</strong> strepsirrhines, but only<br />

the <strong>primate</strong> level correlations survive adjusted<br />

degrees of freedom (Table 5).<br />

Strait (1999) proposed a hypothesis similar<br />

to those of Hofer (1969) <strong>and</strong> Biegert<br />

(1963). Noting that total basi<strong>cranial</strong> length<br />

scales close to isometry with noncortical<br />

brain volume, Strait (1999) suggested that<br />

variation in the midline basi<strong>cranial</strong> angle<br />

might be due to increases in the size of the<br />

telencephalon relative to the noncortical<br />

part of the brain, 4 rather than the size of the<br />

brain relative to the <strong>cranial</strong> <strong>base</strong>. Analysis<br />

of our data set confirms this hypothesis:<br />

there are significant correlations between<br />

flexion <strong>and</strong> the size of the telencephalon<br />

relative to the brain stem across <strong>primate</strong>s,<br />

using CBA4 (Table 5). Moreover, the only<br />

significant correlation between <strong>cranial</strong> <strong>base</strong><br />

angle <strong>and</strong> a neural variable among strepsirrhines<br />

is the comparison of CBA4 with the<br />

ratio of telencephalon to brain-stem volume<br />

(Table 5). However, despite relatively high<br />

correlation coefficients for haplorhines, these<br />

correlations do not remain significant with<br />

phylogenetically adjusted degrees of freedom.<br />

This suggests that brain size relative<br />

to basi<strong>cranial</strong> length may be a better explanation<br />

for flexion, because it appears to be<br />

more independent of phylogenetic effects.<br />

Whether basi<strong>cranial</strong> flexion accommodates<br />

increases in telencephalon volume relative<br />

to brain-stem volume, <strong>and</strong>/or increases<br />

in overall endo<strong>cranial</strong> volume<br />

relative to <strong>cranial</strong> <strong>base</strong> length, the end result<br />

is a change in brain shape. <strong>The</strong> enlarged<br />

telencephalon of <strong>primate</strong>s is an outgrowth<br />

of the rostral end of the brain stem<br />

that communicates with the rest of the<br />

brain through the diencephalon at its root.<br />

4 Strait (1999) refers to this as the “noncortical scaling hypothesis;”<br />

however, the telencephlon consists of more than cortex: it<br />

also includes the white matter <strong>and</strong> the basal ganglia. Here we<br />

evaluate the role of relative telencephalon volume in producing<br />

flexion.


D.E. Lieberman et al.]<br />

PRIMATE CRANIAL BASE 137<br />

Increasing the size of the outer cortex of the<br />

telencephalon (the neocortex) while still<br />

connecting to the rest of the brain through<br />

the diencephalon might be expected to generate<br />

a more spheroid shape, regardless of<br />

any <strong>function</strong>al constraints on skull or brain<br />

shape. In other words, the telencephalon<br />

may be spheroidal because of the geometry<br />

of its connections <strong>and</strong> the way it develops,<br />

rather than for any <strong>function</strong>al or adaptive<br />

reason. Alternately, a spheroidal cerebrum<br />

may minimize “wiring length” in the brain,<br />

a potentially important principle of design<br />

in neural architecture (Allman <strong>and</strong> Kaas,<br />

1974; Barlow, 1986; Mitchison, 1991; Cherniak,<br />

1995; Van Essen, 1997). Accordingly, a<br />

spheroid telencephalon may optimize neocortical<br />

wiring lengths as well as minimize<br />

the distance from all points in the cerebrum<br />

to the diencephalon, a structure through<br />

which all connections to the rest of the brain<br />

must pass (Ross <strong>and</strong> Henneberg, 1995). Another<br />

possible advantage of a flexed basicranium<br />

derives from the in vitro experiments<br />

of Demes (1985), showing that the angulation<br />

of the <strong>cranial</strong> <strong>base</strong> in combination with<br />

a spherical neurocranium helps distribute<br />

applied stresses efficiently over a large area<br />

<strong>and</strong> decreases stresses in the anterior <strong>cranial</strong><br />

<strong>base</strong> during loading of the temporom<strong>and</strong>ibular<br />

joint. This interesting model, however,<br />

requires further testing.<br />

Whether the spheroid shape of the telencephalon<br />

is a <strong>function</strong>al adaptation or a<br />

structural consequence of geometry <strong>and</strong><br />

developmental processes remains to be determined.<br />

Nevertheless, the presence of<br />

the cerebellum, <strong>and</strong> ultimately of the<br />

brain stem, prevents caudal expansion of<br />

the telencephalon, making rostral expansion<br />

of the telencephalon the easiest route.<br />

This would cause the especially large human<br />

brain to develop a kink of the kind<br />

measured by Hofer, which in turn may<br />

cause flexion of the basicranium. If this<br />

hypothesis is correct, then some proportion<br />

of the variation in basi<strong>cranial</strong> angle<br />

among <strong>primate</strong>s is caused by intrinsic<br />

changes in brain shape, <strong>and</strong> not the relationship<br />

between the size of the brain <strong>and</strong><br />

the <strong>base</strong> on which it sits.<br />

One caution (noted above) is that ontogenetic<br />

data suggest that the interspecific<br />

variation in <strong>cranial</strong> <strong>base</strong> angle <strong>and</strong> shape<br />

presented above is partially a consequence<br />

of variables other than relative encephalization<br />

or intrinsic brain shape. Ontogenetic<br />

data are useful because they allow one to<br />

examine temporal relationships among predicted<br />

causal factors. <strong>The</strong> human ontogenetic<br />

data provide mixed support for the<br />

hypothesis that <strong>cranial</strong> <strong>base</strong> angulation reflects<br />

relative encephalization. Jeffery<br />

(1999) found no significant relationship between<br />

CBA1 <strong>and</strong> IRE1 during the second<br />

fetal trimester in humans, when brain<br />

growth is especially rapid; but Lieberman<br />

<strong>and</strong> McCarthy (1999) found that the human<br />

<strong>cranial</strong> <strong>base</strong> flexes rapidly during the first 2<br />

postnatal years, when most brain growth<br />

occurs. Why relative brain size in humans<br />

correlates with <strong>cranial</strong> <strong>base</strong> angle after<br />

birth but not before remains to be explained.<br />

In addition, <strong>and</strong> in contrast to humans,<br />

the <strong>cranial</strong> <strong>base</strong> in all nonhuman <strong>primate</strong>s<br />

so far analyzed extends rather than<br />

flexes during the period of postnatal brain<br />

growth, <strong>and</strong> continues to extend throughout<br />

the period of facial growth, after brain<br />

growth has ceased. In Pan, for example, approximately<br />

88% of <strong>cranial</strong> <strong>base</strong> extension<br />

(CBA1) occurs after the brain has reached<br />

95% adult size (Lieberman <strong>and</strong> McCarthy,<br />

1999). Similar results characterize other<br />

genera (e.g., Macaca; Sirianni <strong>and</strong> Swindler,<br />

1985; Schneiderman, 1992).<br />

Ontogenetic data do not disprove the hypothesis<br />

that variation in <strong>cranial</strong> <strong>base</strong> angle<br />

is related to brain size, but instead highlight<br />

the likelihood that the processes which generate<br />

variation in <strong>cranial</strong> <strong>base</strong> angle are<br />

polyphasic <strong>and</strong> multifactorial. Notably, the<br />

ontogenetic data suggest that the tight<br />

structural relationship between the face<br />

<strong>and</strong> the anterior <strong>cranial</strong> <strong>base</strong> (discussed below)<br />

is also an important influence on <strong>cranial</strong><br />

<strong>base</strong> angle. This suggests that a large<br />

proportion of the interspecific variation in<br />

CBA, IRE, <strong>and</strong> other aspects of neural size<br />

<strong>and</strong> shape reported above is explained by<br />

interactions between the brain <strong>and</strong> the <strong>cranial</strong><br />

<strong>base</strong> prior to the end of the neural<br />

growth phase. <strong>The</strong>reafter, other factors (especially<br />

those related to the face) influence<br />

the shape of the <strong>cranial</strong> <strong>base</strong>. One obvious<br />

way to test this hypothesis is to compare the


138 YEARBOOK OF PHYSICAL ANTHROPOLOGY [Vol. 43, 2000<br />

above interspecific analyses of adults with<br />

comparable analyses of infants at the period<br />

when the brain has ceased growing, but before<br />

much of the face has grown.<br />

Brain volume <strong>and</strong> posterior <strong>cranial</strong><br />

fossa shape<br />

Dean <strong>and</strong> Wood (1981, 1982) <strong>and</strong> Aiello<br />

<strong>and</strong> Dean (1990) hypothesized that increases<br />

in cerebellum size correlate with increases<br />

in the size of the posterior <strong>cranial</strong><br />

fossa. This correlation is purportedly a result<br />

of increases in basi<strong>cranial</strong> flexion; <strong>and</strong><br />

by lateral <strong>and</strong> anterior displacement of the<br />

lateral aspects of the petrous pyramids,<br />

which cause the petrous pyramids to be<br />

more coronally oriented in humans than in<br />

great apes. However, Ross <strong>and</strong> Ravosa<br />

(1993) found little support for a link between<br />

absolute cerebellum volume <strong>and</strong><br />

CBA4; in addition, Spoor (1997) did not find<br />

a correlation between cerebellum volume<br />

<strong>and</strong> petrous orientation. Rather, Spoor<br />

(1997) showed that more coronally oriented<br />

petrous pyramids in adult <strong>primate</strong>s correlate<br />

better with increases in brain volume<br />

relative to basi<strong>cranial</strong> length. In addition,<br />

the petrous pyramids, when viewed from<br />

the internal aspect of the <strong>cranial</strong> <strong>base</strong>, are<br />

not more coronally oriented in humans vs.<br />

other apes (Spoor, 1997).<br />

<strong>The</strong> probable explanation for these results<br />

may be that the posterior <strong>cranial</strong> <strong>base</strong><br />

(Ba-S) scales with isometry against both<br />

cerebellum volume <strong>and</strong> the volume of infratentorial<br />

neural structures (cerebellum, medulla<br />

oblongata, mesencephalon) (Table 4).<br />

If this scaling pattern characterizes those<br />

dimensions of the posterior <strong>cranial</strong> fossa not<br />

in the mid-sagittal plane, then increased<br />

infratentorial neural volumes would not necessitate<br />

changes in the proportions of the<br />

entire posterior <strong>cranial</strong> fossa. This would<br />

imply that variation in the orientation of<br />

the petrous pyramids may be linked to<br />

changes in other <strong>cranial</strong> systems <strong>and</strong> is not<br />

solely a structural response to changes in<br />

cerebellar volume.<br />

ASSOCIATIONS BETWEEN CRANIAL<br />

BASE AND FACE<br />

It has long been known that the <strong>cranial</strong><br />

<strong>base</strong> plays an important role in facial<br />

growth, but many details of how these regions<br />

interact remain poorly understood.<br />

While the face has some influence on <strong>cranial</strong><br />

<strong>base</strong> growth (see below), there are two major<br />

reasons to believe that the <strong>cranial</strong> <strong>base</strong><br />

exerts a greater influence on the face than<br />

vice versa during growth by setting up certain<br />

key spatial relationships. First, the majority<br />

of the <strong>cranial</strong> <strong>base</strong> (with the exception<br />

of the ethmoidal portions of the ethmomaxillary<br />

complex) attains adult size long before<br />

the face (Moore <strong>and</strong> Lavelle, 1974). Second,<br />

as noted above, most of the face grows anteriorly,<br />

laterally, inferiorly, <strong>and</strong> around the<br />

<strong>cranial</strong> <strong>base</strong>. In all mammals, the upper<br />

portion of the face (the orbital <strong>and</strong> upper<br />

nasal regions) grows antero inferiorly relative<br />

to the anterior <strong>cranial</strong> <strong>base</strong> <strong>and</strong> floor;<br />

<strong>and</strong> the middle face (mostly the nasal region)<br />

grows anteriorly relative to the middle<br />

<strong>cranial</strong> fossa. <strong>The</strong> lower portion of the face<br />

(the m<strong>and</strong>ibular <strong>and</strong> maxillary arches <strong>and</strong><br />

their supporting structures) interacts only<br />

indirectly with the <strong>cranial</strong> <strong>base</strong>, since the<br />

maxillary arch grows inferiorly from the<br />

middle face <strong>and</strong> anteriorly relative to the<br />

pterygoid processes of the sphenoid.<br />

<strong>The</strong>se spatial <strong>and</strong> developmental associations<br />

raise an important question: to what<br />

extent does the <strong>cranial</strong> <strong>base</strong> influence facial<br />

growth <strong>and</strong> form? In order to address this<br />

issue, we first discuss the relationships between<br />

two regions of the face <strong>and</strong> <strong>cranial</strong><br />

<strong>base</strong> that are contiguous across <strong>function</strong>al<br />

or developmental boundaries (the so-called<br />

growth counterparts of Enlow, 1990): 1) the<br />

anterior <strong>cranial</strong> fossa <strong>and</strong> the upper, orbital,<br />

<strong>and</strong> nasal portions of the face, <strong>and</strong> 2) the<br />

middle <strong>cranial</strong> fossa <strong>and</strong> the middle, ethmomaxillary<br />

portion of the face. We conclude<br />

with a brief discussion of the possible relationships<br />

between <strong>cranial</strong> <strong>base</strong> shape <strong>and</strong><br />

overall facial shape.<br />

Anterior <strong>cranial</strong> fossa shape <strong>and</strong> upper<br />

facial growth<br />

<strong>The</strong> upper face comprises the orbital cavities,<br />

the orbital superstructures, <strong>and</strong> the<br />

upper portion of the nasal cavity. <strong>The</strong> upper<br />

face therefore incorporates elements of the<br />

anterior <strong>cranial</strong> <strong>base</strong>, including the ethmoid,<br />

parts of the sphenoid, <strong>and</strong> significant<br />

portions of the frontal bone. <strong>The</strong> upper face


D.E. Lieberman et al.]<br />

PRIMATE CRANIAL BASE 139<br />

grows away from the rest of the <strong>cranial</strong> <strong>base</strong><br />

in three ways. Initially, as the eyeballs exp<strong>and</strong>,<br />

the orbital cavity exp<strong>and</strong>s anteriorly,<br />

inferiorly, <strong>and</strong> laterally through drift <strong>and</strong><br />

displacement (Moss <strong>and</strong> Young, 1960; Enlow,<br />

1990). Animals enucleated during the<br />

period of eyeball growth consequently have<br />

deficient anterior <strong>and</strong> lateral growth of the<br />

upper face (see Sarnat, 1982). In addition,<br />

since the roof of the orbit also contributes to<br />

the floor of the anterior <strong>cranial</strong> fossa (i.e.,<br />

the orbital plates of the frontal bone <strong>and</strong><br />

lesser wings of the sphenoid), the position,<br />

orientation, <strong>and</strong> shape of the orbital roof<br />

must inevitably be affected by growth of the<br />

frontal lobes <strong>and</strong> anterior <strong>cranial</strong> fossa. Finally,<br />

the orbital cavities <strong>and</strong> superstructures<br />

grow anteriorly <strong>and</strong> laterally away<br />

from the anterior <strong>cranial</strong> fossa, but only after<br />

some period of postnatal development.<br />

In humans, for example, the upper face does<br />

not emerge from under the anterior <strong>cranial</strong><br />

<strong>base</strong> until after the eruption of the second<br />

molars (Riolo et al., 1974; Lieberman, 2000).<br />

In contrast, the front of the upper face in<br />

most nonhuman <strong>primate</strong>s projects anteriorly<br />

relative to the front of the anterior <strong>cranial</strong><br />

fossa prior to the eruption of the first<br />

molars (Krogman, 1969; Lieberman, 1998,<br />

2000). Variation in this spatial separation,<br />

termed neuro<strong>cranial</strong>-orbital or neuro-orbital<br />

disjunction, has been analyzed by various<br />

researchers to explain ontogenetic <strong>and</strong><br />

interspecific patterns of supraorbital torus<br />

morphology (Weidenreich, 1941; Moss <strong>and</strong><br />

Young, 1960; Radinsky, 1968, 1970, 1977,<br />

1979; Shea, 1985a, 1986, 1988; Ravosa,<br />

1988, 1991a,b; Hyl<strong>and</strong>er <strong>and</strong> Ravosa, 1992;<br />

Vinyard <strong>and</strong> Smith, 1997; Lieberman, 1998,<br />

2000; Ravosa et al., 2000b).<br />

<strong>The</strong>se developmental <strong>and</strong> architectural<br />

relationships between the anterior <strong>cranial</strong><br />

<strong>base</strong> <strong>and</strong> upper face have several consequences<br />

for interspecific variation in <strong>primate</strong>s.<br />

First, the orientation of the orbits<br />

<strong>and</strong> the upper face is intimately related to<br />

the orientation of the anterior <strong>cranial</strong> <strong>base</strong><br />

in haplorhines (Ravosa, 1991a,b; Ross <strong>and</strong><br />

Ravosa, 1993; Ross <strong>and</strong> Henneberg, 1995),<br />

confirming the hypothesis of Dabelow (1929,<br />

1931) that the orientation of the orbits <strong>and</strong><br />

the anterior <strong>cranial</strong> <strong>base</strong> should be tightly<br />

correlated in animals like haplorhines,<br />

whose orbits are closely approximated below<br />

the olfactory tract (see also Cartmill,<br />

1970, 1972; Ravosa et al., 2000a,2000b).<br />

Several recent studies have found support<br />

for this hypothesis. First, brow-ridge size in<br />

anthropoids (but not strepsirhines) is highly<br />

correlated with variation in the position of<br />

the orbits relative to the anterior <strong>cranial</strong><br />

<strong>base</strong> (Ravosa, 1988, 1991a,b; Hyl<strong>and</strong>er <strong>and</strong><br />

Ravosa, 1992; Vinyard <strong>and</strong> Smith, 1997;<br />

Lieberman, 2000; Ravosa et al., 2000b). Finally,<br />

there is evidence that the upper face<br />

tends to be rotated dorsally relative to the<br />

posterior <strong>cranial</strong> <strong>base</strong> in <strong>primate</strong>s with a<br />

highly extended <strong>cranial</strong> <strong>base</strong>, but more ventrally<br />

rotated relative to the posterior <strong>cranial</strong><br />

<strong>base</strong> in <strong>primate</strong>s such as humans <strong>and</strong><br />

bonobos with a more flexed <strong>cranial</strong> <strong>base</strong> (Delattre<br />

<strong>and</strong> Fenart, 1956; Moss <strong>and</strong> Young,<br />

1960; Heintz, 1966; Cramer, 1977; Shea,<br />

1985a, 1986, 1988; Ravosa, 1988, 1991a,b;<br />

Lieberman, 2000).<br />

While the structural boundaries shared<br />

by the anterior <strong>cranial</strong> <strong>base</strong> <strong>and</strong> the upper<br />

face result in a high degree of integration<br />

between these two regions, the extent to<br />

which anterior <strong>cranial</strong> <strong>base</strong> shape influences<br />

other aspects of facial shape is less<br />

clear. In an interspecific study of 68 species,<br />

Ross <strong>and</strong> Ravosa (1993) found that both orbit<br />

<strong>and</strong> palate orientation are significantly<br />

correlated with anterior <strong>cranial</strong> <strong>base</strong> orientation,<br />

but that palate orientation accounts<br />

for none of the variation in <strong>cranial</strong> <strong>base</strong><br />

angle independent of orbital axis orientation.<br />

McCarthy <strong>and</strong> Lieberman (2001) have<br />

also shown that the orientation of the orbits<br />

<strong>and</strong> the anterior <strong>cranial</strong> <strong>base</strong> are correlated<br />

with each other (r 0.617, P 0.05) in<br />

haplorhines but not in strepsirhines. <strong>The</strong>se<br />

studies suggest that the orientation of the<br />

anterior <strong>cranial</strong> <strong>base</strong> affects the orientation<br />

of the upper face directly, but that it only<br />

indirectly influences palate orientation<br />

through the integration of palate <strong>and</strong> orbits<br />

(see also Ravosa, 1988; Ravosa <strong>and</strong> Shea,<br />

1994). Consequently, it seems likely that<br />

the anterior <strong>cranial</strong> <strong>base</strong> exerts a slight influence<br />

on facial orientation as a whole, but<br />

that only the orbital region of the face is<br />

directly integrated with the anterior <strong>cranial</strong><br />

<strong>base</strong>. Higher levels of integration appear to<br />

characterize those organisms with greater


140 YEARBOOK OF PHYSICAL ANTHROPOLOGY [Vol. 43, 2000<br />

TABLE 6. Descriptive statistics for orientation of the palate relative to the anterior <strong>cranial</strong> <strong>base</strong> in <strong>primate</strong>s,<br />

haplorhini, <strong>and</strong> strepsirrhini 1<br />

Planum sphenoideum<br />

Line from Sella to foramen caecum<br />

N Mean SD Minimum Maximum N Mean SD Minimum Maximum<br />

Primates 63 15.1 15.66 41.5 45.2 60 20.0 15.15 22.1 67.3<br />

Haplorhines 47 12.80 16.35 41.5 45.2 46 15.1 12.14 22.1 52.0<br />

Strepsirrhines 16 21.8 11.37 1.7 42.3 14 36.1 12.84 19.8 67.3<br />

1 Calculated from data in Ross <strong>and</strong> Ravosa (1993) <strong>and</strong> Ross <strong>and</strong> Henneberg (1995), <strong>and</strong> presented above by subtracting AFK from<br />

CBA4 <strong>and</strong> CBA1.<br />

encephalization, increased orbital convergence,<br />

<strong>and</strong> relatively large orbits (Ravosa et<br />

al., 2000a).<br />

Enlow (1990) proposed several structural<br />

relationships between the anterior <strong>cranial</strong><br />

<strong>base</strong> <strong>and</strong> various aspects of facial shape. His<br />

general model for the plan of the face is<br />

<strong>base</strong>d almost exclusively on analyses of human<br />

radiographs, <strong>and</strong> has not been carefully<br />

tested in most respects. However,<br />

three hypotheses are of special interest.<br />

First, since the floor of the nasal chamber is<br />

also the roof of the oral cavity, Enlow (1990)<br />

proposed that interorbital breadth should<br />

be correlated with prognathism because<br />

“the broad nasal <strong>base</strong> of most other mammals<br />

supports a correspondingly much<br />

longer snout.” Smith <strong>and</strong> Josell (1984)<br />

tested this hypothesis using a sample of 32<br />

<strong>primate</strong>s, <strong>and</strong> found a low correlation (r <br />

0.46) between interorbital breadth <strong>and</strong><br />

m<strong>and</strong>ibular prognathism, <strong>and</strong> noted that<br />

the correlation was even weaker when effects<br />

of overall <strong>cranial</strong> length <strong>and</strong> body size<br />

were taken into account. This is perhaps not<br />

surprising, as there is no obvious biomechanical<br />

or developmental reason why the<br />

length of the face should be correlated with<br />

the width of the posterior part of the face.<br />

Second, Enlow (1990) proposed that the orientation<br />

of the cribriform plate in all mammals,<br />

including <strong>primate</strong>s, is perpendicular<br />

to the orientation of the nasomaxillary complex<br />

(defined as a plane from the most anterior<br />

point on the frontal squama to the<br />

prosthion). Ravosa <strong>and</strong> Shea (1994) investigated<br />

the angle between the cribriform<br />

plane (measured in two ways) <strong>and</strong> the orientation<br />

of the midface (measured from the<br />

prosthion tangent to the endo<strong>cranial</strong> contour<br />

of the frontal) in a sample of Old World<br />

monkeys. <strong>The</strong>y found that when the cribriform<br />

plane was defined as the line connecting<br />

the anterior-most <strong>and</strong> posterior-most aspects<br />

of the cribriform plate, the mean angle<br />

between these planes was 93° for cercopithecines<br />

<strong>and</strong> 86° for colobines <strong>and</strong> was<br />

independent of face size. However, variability<br />

in this angle was relatively high; <strong>and</strong><br />

when the cribriform angle was measured as<br />

the line connecting the anterior-most <strong>and</strong><br />

posterior-most aspects of the cribriform<br />

plate at the ethmoid/nasal cavity junction,<br />

the angle was significantly different from<br />

90° <strong>and</strong> correlated with skull size. Thus,<br />

there is only limited support for the hypothesis<br />

that the cribriform plate <strong>and</strong> the orientation<br />

of the snout are perpendicular. Finally,<br />

Enlow (1990) <strong>and</strong> several others (e.g.,<br />

Sirianni <strong>and</strong> Swindler, 1979) have suggested<br />

that the palatal plane <strong>and</strong> the anterior<br />

<strong>cranial</strong> <strong>base</strong> (S-FC) should be roughly<br />

parallel across <strong>primate</strong>s when the cribriform<br />

plate is parallel to the planum sphenoideum.<br />

However, interspecific analysis of<br />

the orientation of the palate relative to both<br />

the planum sphenoideum <strong>and</strong> a line from<br />

the sella to the foramen caecum (Table 6)<br />

provide little support for the hypothesis of<br />

Enlow (1990) across Primates, Haplorhini,<br />

<strong>and</strong> Strepsirhini.<br />

Middle <strong>cranial</strong> fossa shape <strong>and</strong><br />

midfacial growth<br />

<strong>The</strong> second major region of interaction between<br />

the <strong>cranial</strong> <strong>base</strong> <strong>and</strong> the face is along<br />

the junction between the middle <strong>cranial</strong><br />

fossa <strong>and</strong> the posterior margins of the midface.<br />

Whereas the upper face <strong>and</strong> anterior<br />

<strong>cranial</strong> fossa are tightly integrated because<br />

they share many of the same bony walls, the<br />

middle portion of the face (the ethmomaxillary<br />

complex) <strong>and</strong> middle <strong>cranial</strong> fossa are<br />

growth counterparts that may interact<br />

along their complex boundary, which lies<br />

more or less in the coronal plane (Hoyte,


D.E. Lieberman et al.]<br />

PRIMATE CRANIAL BASE 141<br />

Fig. 10. Posterior maxillary (PM) plane <strong>and</strong> 90° orientation relative to the neutral horizontal axis<br />

(NHA) of the orbits; illustrated here in, Homo sapiens. <strong>The</strong> inferior <strong>and</strong> superior termini of the PM plane<br />

are, respectively, pterygomaxillary (Ptm) <strong>and</strong> the PM point (PMp). <strong>The</strong> anterior <strong>and</strong> posterior termini of<br />

the NHA are, respectively, OM <strong>and</strong> OA (see text for definitions).<br />

1991). In particular, the ethmomaxillary<br />

complex grows anteriorly, laterally, <strong>and</strong> inferiorly<br />

away from the middle <strong>cranial</strong> fossa<br />

at a number of primary growth sites (e.g.,<br />

the spheno-palatine suture, the spheno-zygomatic<br />

suture, <strong>and</strong> the spheno-ethmoid<br />

synchondrosis). Consequently, the shape of<br />

the middle <strong>cranial</strong> fossa, especially the<br />

greater wings of the sphenoid (which house<br />

the temporal lobes), must also play some<br />

role in influencing the orientation of the<br />

posterior margin of the ethmomaxillary<br />

complex <strong>and</strong> its position relative to the rest<br />

of the <strong>cranial</strong> <strong>base</strong>.<br />

Recent research on the integration of the<br />

middle <strong>cranial</strong> fossa <strong>and</strong> the midface has<br />

focused on the role of the posterior maxillary<br />

(PM) plane. <strong>The</strong> PM plane has been<br />

defined in several different ways (Enlow<br />

<strong>and</strong> Azuma, 1975; Enlow, 1976, 1990; Enlow<br />

<strong>and</strong> Hans, 1996), but we use here the<br />

definition of Enlow <strong>and</strong> Azuma (1975), as<br />

the line connecting two termini: 1) pterygomaxillary<br />

(Ptm), the average midline point<br />

of the most inferior <strong>and</strong> posterior points on<br />

the maxillary tuberosities; <strong>and</strong> 2) the PM<br />

point (PMp), the average midline point of<br />

the anterior-most points on the lamina of<br />

the greater wings of the sphenoid (for details,<br />

see McCarthy <strong>and</strong> Lieberman, 2001).<br />

<strong>The</strong>se points <strong>and</strong> their relationship to the<br />

<strong>cranial</strong> <strong>base</strong> <strong>and</strong> face are illustrated in Figure<br />

10. Note that because the PM plane is<br />

defined using two paired registration<br />

points, it is technically not a plane, but is<br />

instead an abstract line whose termini do<br />

not lie in the same parasagittal plane. Despite<br />

its abstract nature, the PM plane is a


142 YEARBOOK OF PHYSICAL ANTHROPOLOGY [Vol. 43, 2000<br />

potentially useful analytical concept for researchers<br />

interested in integration between<br />

the <strong>cranial</strong> <strong>base</strong> <strong>and</strong> face because it effectively<br />

characterizes both the posterior margin<br />

of the face <strong>and</strong> the boundary between<br />

the anterior <strong>and</strong> middle <strong>cranial</strong> fossae in<br />

lateral radiographs. <strong>The</strong> inferior terminus,<br />

Ptm, is the posterolateral corner of the ethmomaxillary<br />

complex <strong>and</strong> lies just in front<br />

of the spheno-palatine suture (Williams et<br />

al., 1995). <strong>The</strong> superior terminus, PMp, is<br />

the anterior-most point of the middle <strong>cranial</strong><br />

fossa, lying close to the midpoint of the<br />

spheno-ethmoid synchondrosis <strong>and</strong> the midpoint<br />

of the spheno-frontal suture on the<br />

floor of the <strong>cranial</strong> <strong>base</strong> in all <strong>primate</strong>s (Van<br />

der Linden <strong>and</strong> Enlow, 1971; McCarthy,<br />

2001; McCarthy <strong>and</strong> Lieberman, 2001).<br />

Perhaps the most interesting aspect of the<br />

PM plane is its relationship to the orbits<br />

<strong>and</strong> the anterior <strong>cranial</strong> <strong>base</strong>. Several researchers<br />

have claimed that the PM plane<br />

always forms a 90° angle to the neutral horizontal<br />

axis (NHA) of the orbits (see Measurement<br />

Definitions). In their initial study,<br />

Enlow <strong>and</strong> Azuma (1975) found the PM-<br />

NHA angle to average 90° in a combined<br />

mammalian sample of 45 species, <strong>and</strong> 90° in<br />

a large sample of adult humans. Ravosa<br />

(1991a,b), <strong>and</strong> Ravosa <strong>and</strong> Shea (1994)<br />

tested the PM-NHA angle in a cross-sectional<br />

sample of macaques <strong>and</strong> two interspecific<br />

sample of adult <strong>primate</strong>s, <strong>and</strong> obtained<br />

consistent, but different PM-NHA<br />

angles from those of Enlow <strong>and</strong> Azuma<br />

(1975), that ranged between 18° <strong>and</strong> 5° below<br />

90°. However, these studies measured<br />

the PM plane <strong>and</strong> the NHA (the latter only<br />

slightly) differently, <strong>and</strong> several more recent<br />

studies have corroborated the original<br />

hypothesis of Enlow <strong>and</strong> Azuma (1975). In<br />

particular, Bromage (1992) found the PM-<br />

NHA angle in a cross-sectional sample of 45<br />

Pan troglodytes crania to be 89.2 3.4° SD<br />

for dental stage I, 90.5 3.1° SD for dental<br />

stage II, <strong>and</strong> 88.2 4.0° SD for dental stage<br />

III. However, these data show some significant<br />

variation during growth, <strong>and</strong> some<br />

adult crania have PM-NHA angles somewhat<br />

different from 90°, especially those for<br />

certain hominids. Lieberman (1998) found<br />

the PM-NHA angle to be 89.9 1.7° SD in a<br />

longitudinal series of humans (Denver<br />

Fig. 11. Histograms comparing mean PM-NHA angle<br />

in samples of 18 adult anthropoid species (top) <strong>and</strong><br />

15 adult strepsirrhines species (bottom). PM-NHA° is<br />

not significantly different from 90° in any species.<br />

Growth Study; n 353) aged 1 month<br />

through 17 years, 9 months. Also, McCarthy<br />

<strong>and</strong> Lieberman (in press) recently found the<br />

PM-NHA angle to average 90.0 0.38° SD<br />

in a pooled sample of adults from 18 anthropoid<br />

species, <strong>and</strong> 89.4 0.46° SD in a pooled<br />

sample of adults from 15 strepsirhine species<br />

(Fig. 11). Consequently, the PM-NHA<br />

does appear to be invariant in <strong>primate</strong>s,<br />

with values for the most part near 90°. It<br />

should be stressed, however, that the developmental<br />

<strong>and</strong> <strong>function</strong>al <strong>base</strong>s (if any) for<br />

this purported invariance are still unknown<br />

<strong>and</strong> require further study.<br />

<strong>The</strong> 90° PM-NHA angle is useful for examining<br />

craniofacial integration <strong>and</strong> variation<br />

because, as noted above, the NHA is<br />

tightly linked to the orientation of the anterior<br />

<strong>cranial</strong> fossa <strong>and</strong> the ethmomaxillary<br />

complex. <strong>The</strong> roofs of the orbits (which help


D.E. Lieberman et al.]<br />

PRIMATE CRANIAL BASE 143<br />

define the NHA) comprise much of the floor<br />

of the anterior <strong>cranial</strong> fossa. <strong>The</strong>refore, it<br />

follows that the PM plane <strong>and</strong> the anterior<br />

<strong>cranial</strong> <strong>base</strong> should also form an approximately<br />

90° angle in <strong>primate</strong>s whose orbits<br />

are approximated to the midline. This hypothesis<br />

was tested by McCarthy <strong>and</strong><br />

Lieberman (2001), who found that the angle<br />

between the PM plane <strong>and</strong> the planum<br />

sphenoideum averaged 95.2 7.6° SD (n <br />

18) in anthropoids <strong>and</strong> 82.8 9.5° SD (n <br />

14) in strepsirhines. McCarthy <strong>and</strong> Lieberman<br />

(2001) also found that the angle between<br />

the PM plane <strong>and</strong> the midline anterior<br />

<strong>cranial</strong> <strong>base</strong> (from the sella to the<br />

foramen caecum) averages 89.2 9.97° SD<br />

(n 18) in anthropoids, but 70.6 10.5° SD<br />

(n 15) in strepsirrhines. <strong>The</strong> high st<strong>and</strong>ard<br />

deviations of these angles indicate that<br />

the integration between the back of the face<br />

<strong>and</strong> the anterior <strong>cranial</strong> <strong>base</strong> is not very<br />

strong. Ross <strong>and</strong> Ravosa (1993) also came to<br />

similar conclusions by comparing the orbital<br />

axis orientation relative to the posterior <strong>cranial</strong><br />

<strong>base</strong>, against the orientation of the planum<br />

sphenoideum relative to the posterior<br />

<strong>cranial</strong> <strong>base</strong>. <strong>The</strong> differences between anthropoids<br />

<strong>and</strong> strepsirhines in the relationship<br />

of the orbits to the <strong>cranial</strong> <strong>base</strong> can be<br />

explained by the fact that the roof of the<br />

orbits does not contribute to the midline<br />

<strong>cranial</strong> <strong>base</strong> in strepsirhines, <strong>and</strong> because<br />

the cribriform plate tends to be oriented<br />

more vertically relative to the planum sphenoideum<br />

in strepsirrhines than in anthropoids<br />

(Cartmill, 1970).<br />

<strong>The</strong> potential integration of the middle<br />

<strong>and</strong> anterior <strong>cranial</strong> fossae with the face (as<br />

measured via the PM plane) <strong>and</strong> the anterior<br />

<strong>cranial</strong> <strong>base</strong> raises several interesting<br />

issues. Most importantly, the top <strong>and</strong> back<br />

of the face appear to form an integrated<br />

unit, the “facial block” which rotates during<br />

<strong>ontogeny</strong> around an axis through the intersection<br />

of the anterior <strong>and</strong> middle <strong>cranial</strong><br />

fossae at the front of the greater wings of<br />

the sphenoid (McCarthy <strong>and</strong> Lieberman,<br />

2001). This facial block is characteristic of<br />

anthropoids but not strepsirhines, <strong>and</strong> manifests<br />

itself through correlations between<br />

<strong>cranial</strong> <strong>base</strong> angle <strong>and</strong> upper facial orientation<br />

in <strong>primate</strong>s (Weidenriech, 1941; Moss<br />

<strong>and</strong> Young, 1960; Biegert, 1963; Shea,<br />

1985a, 1986, 1988; Ravosa, 1988, 1991a,b;<br />

Ross <strong>and</strong> Ravosa, 1993; Ross, 1995a,b; May<br />

<strong>and</strong> Sheffer, 1999; Lieberman, 2000; Ravosa<br />

et al., 2000a, 2000b). In particular, as the<br />

anterior <strong>cranial</strong> <strong>base</strong> flexes relative to the<br />

posterior <strong>cranial</strong> <strong>base</strong>, the PM plane also<br />

must flex relative to the posterior <strong>cranial</strong><br />

<strong>base</strong>, rotating the posterior <strong>and</strong> upper portions<br />

of the face underneath the anterior<br />

<strong>cranial</strong> fossa (klinorhynchy). In contrast, extension<br />

of the anterior <strong>cranial</strong> <strong>base</strong> relative<br />

to the posterior <strong>cranial</strong> <strong>base</strong> will rotate the<br />

posterior <strong>and</strong> upper portions of the face dorsally<br />

relative to the posterior <strong>cranial</strong> <strong>base</strong><br />

(airorhynchy) (Fig. 12).<br />

<strong>The</strong> relationship of the orientation of the<br />

back of the face (as measured for example by<br />

the PM plane) to the anterior <strong>cranial</strong> <strong>base</strong><br />

also influences nasopharynx shape. As Figure<br />

12 shows, flexion of the anterior <strong>cranial</strong><br />

<strong>base</strong> <strong>and</strong>/or face relative to the posterior<br />

<strong>cranial</strong> <strong>base</strong> not only rotates the face under<br />

the anterior <strong>cranial</strong> fossa, but it also shortens<br />

(absolutely <strong>and</strong> relatively) the length of<br />

the pharyngeal space between the back of<br />

the palate <strong>and</strong> the front of the vertebral<br />

column (Laitman <strong>and</strong> Heimbuch, 1982;<br />

Spoor et al., 1999; McCarthy <strong>and</strong> Lieberman,<br />

2001). While flexion of the <strong>cranial</strong> <strong>base</strong><br />

during <strong>ontogeny</strong> is completely independent<br />

of the descent of the hyoid <strong>and</strong> larynx<br />

(Lieberman <strong>and</strong> McCarthy, 1999), variation<br />

in <strong>cranial</strong> <strong>base</strong> angle does influence some<br />

aspects of pharyngeal shape (Laitman <strong>and</strong><br />

Heimbuch, 1982; see below).<br />

Ross <strong>and</strong> Henneberg (1995) suggested<br />

that there must be <strong>function</strong>al constraints on<br />

how far back the hard palate can be positioned<br />

without occluding the airway. <strong>The</strong><br />

integration of the anterior <strong>cranial</strong> <strong>base</strong> with<br />

the upper <strong>and</strong> posterior margins of the face<br />

means that these constraints on pharynx<br />

position might determine the maximum<br />

possible degree of basi<strong>cranial</strong> angle, particularly<br />

in genera such as Pongo <strong>and</strong> Alouatta<br />

with relatively large pharyngeal structures<br />

(Biegert, 1957, 1963). Ross <strong>and</strong> Henneberg<br />

(1995) suggested that hominoids might<br />

have found a way to circumvent these “constraints.”<br />

Hominoids have more airorhynch<br />

(dorsally rotated <strong>and</strong> less frontated) orbits<br />

<strong>and</strong> palates than nonhominoid <strong>primate</strong>s<br />

with comparably flexed basicrania (Shea,


144 YEARBOOK OF PHYSICAL ANTHROPOLOGY [Vol. 43, 2000<br />

Fig. 12. Proposed “facial block,” showing the effects of angular invariance between the back of the face<br />

(summarized by the PM plane) <strong>and</strong> the top of the face, which is also the bottom of the anterior <strong>cranial</strong><br />

<strong>base</strong> (S-FC). Changes in <strong>cranial</strong> <strong>base</strong> angle cause the top <strong>and</strong> back of the face to rotate together around<br />

an imaginary axis through the PM point. See text for details.<br />

1985a, 1988; Ross <strong>and</strong> Henneberg, 1995).<br />

This airorhynchy has yet to be explained<br />

developmentally <strong>and</strong> <strong>function</strong>ally; however,<br />

Ross <strong>and</strong> Henneberg (1995) suggested that<br />

it evolved in hominoids in response to increased<br />

flexion of the <strong>cranial</strong> <strong>base</strong> producing<br />

posterior displacement of the palate.<br />

More research is needed on the integration<br />

of the midface <strong>and</strong> <strong>cranial</strong> <strong>base</strong>. In particular,<br />

why is the PM plane oriented at 90°<br />

relative to the NHA during postnatal <strong>ontogeny</strong><br />

<strong>and</strong> thus across taxa? Another question<br />

of interest is, what aspects of <strong>cranial</strong> <strong>base</strong><br />

<strong>and</strong> facial shape are responsible for most of<br />

the variation in PM plane position, <strong>and</strong><br />

hence facial orientation? This problem has<br />

not been well studied, but the orientation of<br />

the PM plane is probably most affected by<br />

the size of the middle <strong>cranial</strong> fossa, especially<br />

the length of the temporal lobes, by<br />

flexion of the sphenoid, <strong>and</strong> by the length of<br />

the anterior sphenoid in the midline <strong>cranial</strong><br />

<strong>base</strong>. Three-dimensional studies of the interface<br />

between the PM plane <strong>and</strong> the <strong>cranial</strong><br />

<strong>base</strong> are needed to resolve these <strong>and</strong><br />

other questions about <strong>cranial</strong> <strong>base</strong>-face interrelations<br />

<strong>and</strong> interactions.<br />

Basi<strong>cranial</strong> width <strong>and</strong> overall facial<br />

shape in humans<br />

Although it is clear that the <strong>cranial</strong> <strong>base</strong><br />

plays a major role in influencing facial orientation<br />

relative to the neurocranium, there<br />

is less information about the potential influence<br />

of the <strong>cranial</strong> <strong>base</strong> on other aspects of<br />

facial shape such as height, length, <strong>and</strong><br />

width. To what extent is overall facial shape<br />

independent of the <strong>cranial</strong> <strong>base</strong>? It is commonly<br />

assumed that the majority of facial<br />

growth is independent of <strong>cranial</strong> <strong>base</strong><br />

growth, largely because much of the face<br />

grows in a skeletal growth trajectory after<br />

the end of the neural growth phase. In humans,<br />

for example, the face attains 95%<br />

adult size by 16–18 years, at least 10 years<br />

after the <strong>cranial</strong> <strong>base</strong> reaches adult size<br />

(Stamrud, 1959; Moore <strong>and</strong> Lavelle, 1974).<br />

In addition, most facial <strong>and</strong> basi<strong>cranial</strong> dimensions<br />

appear to be genetically independent<br />

in adults (Cheverud, 1996). However,<br />

there is some evidence to suggest that<br />

changes in the proportions of the <strong>cranial</strong><br />

<strong>base</strong> can influence facial shape. This interaction<br />

is predicted to be especially important,<br />

<strong>and</strong> perhaps exclusive to humans, in<br />

which the upper face lies almost completely<br />

underneath the anterior <strong>cranial</strong> fossa (Weidenreich,<br />

1941; Howells, 1973; Enlow <strong>and</strong><br />

Bhatt, 1984; Enlow, 1990; Lieberman et al.,<br />

2000).<br />

<strong>The</strong> most explicit of these hypotheses is<br />

that of Enlow (1990), who suggested that<br />

humans with absolutely narrow <strong>cranial</strong><br />

<strong>base</strong>s (primarily dolichocephalics) tend to


D.E. Lieberman et al.]<br />

PRIMATE CRANIAL BASE 145<br />

Fig. 13. Diagram illustrating angles <strong>and</strong> lines used<br />

to calculate foramen magnum orientation relative to<br />

orbital axis orientation <strong>and</strong> head-neck angle (Strait <strong>and</strong><br />

Ross, 1999). <strong>The</strong> orientation of the orbital aperture <strong>and</strong><br />

the neck inclination were measured relative to gravity<br />

from videos (see Strait <strong>and</strong> Ross, 1999, for details). <strong>The</strong><br />

orientation of the foramen magnum relative to the orbital<br />

axis was calculated from measures of the orientation<br />

of both these planes relative to the clivus ossis<br />

occipitalis (CO) as 180°-AOA-FM CO.<br />

have longer, more flexed <strong>cranial</strong> <strong>base</strong>s, <strong>and</strong><br />

narrower faces than individuals with absolutely<br />

wider <strong>cranial</strong> <strong>base</strong>s (primarily<br />

brachycephalics). If variation in overall facial<br />

size is partially independent of <strong>cranial</strong><br />

<strong>base</strong> <strong>and</strong> neuro<strong>cranial</strong> size, then interactions<br />

between <strong>cranial</strong> <strong>base</strong> width <strong>and</strong> facial<br />

width may have some effects on facial<br />

height <strong>and</strong> length. Enlow (1990) proposed<br />

that humans with absolutely narrower <strong>cranial</strong><br />

<strong>base</strong>s tend to have proportionately narrower<br />

<strong>and</strong> antero-posteriorly longer faces<br />

(leptoproscopy) than humans with wider<br />

<strong>cranial</strong> <strong>base</strong>s, who tend to have proportionately<br />

wider <strong>and</strong> antero-posteriorly shorter<br />

faces (euryproscopy). <strong>The</strong> hypothesis receives<br />

some support from studies of artificial<br />

<strong>cranial</strong> deformation in humans. Antón<br />

(1989, 1994), for example, showed that antero-posterior<br />

head-binding during the first<br />

years of life causes not only a wider neurocranium<br />

but also a concomitantly wider face<br />

from additional growth in the most lateral<br />

regions; conversely, circumferential headbinding<br />

results in a narrower neurocranium<br />

<strong>and</strong> face. Lieberman et al. (2000) attempted<br />

to test the hypothesis of Enlow (1990) more<br />

directly with a partial correlation analysis<br />

of a sample of 98 adults from five geographically<br />

<strong>and</strong> craniometrically diverse populations.<br />

<strong>The</strong> study, however, found a low correlation<br />

between upper facial breadth <strong>and</strong><br />

maximum anterior <strong>cranial</strong> fossa breadth<br />

(r 0.53, P 0.001), <strong>and</strong> between midfacial<br />

breadth <strong>and</strong> maximum middle <strong>cranial</strong> fossa<br />

breadth (r 0.49; P 0.001) when differences<br />

in overall size were accounted for using<br />

partial correlation analysis. In addition,<br />

there was only a low tendency (r 0.49, P <br />

0.001) 5 for individuals with narrow <strong>cranial</strong><br />

<strong>base</strong>s to have longer, narrower faces than<br />

individuals with wider <strong>cranial</strong> <strong>base</strong>s, <strong>and</strong><br />

the trend may largely be a factor of interpopulation<br />

rather than intrapopulation<br />

variation. Further studies are needed to<br />

better underst<strong>and</strong> these sources of varia-<br />

5 r 0.40, P 0.001 when facial size is held constant (for<br />

details, see Lieberman et al., 2000).


TABLE 7. Mean values for measures of <strong>cranial</strong> <strong>base</strong> angulation, foramen magnum orientation, <strong>cranial</strong> <strong>base</strong> length, <strong>and</strong> measures of endo<strong>cranial</strong> volume,<br />

including data from Ross <strong>and</strong> Ravosa (1993), Ross <strong>and</strong> Henneberg (1995), <strong>and</strong> new data collected for this study<br />

Species or specimen IRE1 1 IRE5 2 (degrees)<br />

CBA4 1<br />

CBA1<br />

(degrees)<br />

AOA 1<br />

(degrees)<br />

AFK 1<br />

(degrees)<br />

Op-Ba clivus<br />

(degrees) 2<br />

Basion-sella<br />

(mm) 2<br />

Sella-foramen<br />

cecum (mm) 2<br />

Alouatta belzebul 0.680 0.062 186.0 183.3 192.670 175.170 116.170 23.223 36.344<br />

Alouatta palliata 0.710 0.065 188.0 180.7 190.170 174.830 23.509 34.252<br />

Aotus trivirgatus 0.901 0.082 180.0 176.8 168.500 147.830 126.500 11.250 21.014<br />

Arctocebus calabarensis 0.761 0.063 168.0 194.4 175.800 157.200 128.200 13.100 17.320<br />

Ateles fusciceps 0.916 0.085 162.0 170.3 169.500 156.660 118.830 22.593 33.015<br />

Avahi laniger 0.824 0.068 180.0 183.7 174.830 153.170 123.500 13.625 18.224<br />

Brachyteles arachnoides 0.883 0.083 161.0 172.5 170.400 159.400 126.500 24.024 34.352<br />

Cacajao calvus 0.860 0.087 172.0 168.2 173.000 158.830 122.000 19.757 27.689<br />

Callicebus moloch 0.775 0.074 176.0 176.0 164.000 156.400 131.800 15.320 19.631<br />

Callimico goeldi 0.817 0.078 171.0 165.0 154.670 141.500 133.200 11.175 18.425<br />

Callithrix argentata 0.771 0.074 164.0 164.8 152.000 144.330 123.830 10.443 16.929<br />

Cebuella pygmaea 0.880 0.071 171.0 164.2 153.000 141.500 115.330 8.454 13.906<br />

Cebus apella 0.868 0.082 173.0 166.6 165.830 154.170 123.000 20.794 27.890<br />

Cercocebus torquatus 0.874 0.083 172.0 157.3 161.400 142.000 130.000 21.734 33.663<br />

Cercopithecus aethiops 0.881 0.084 171.0 165.2 163.600 143.400 128.400 18.951 28.462<br />

Cercopithecus campbelli 0.879 0.079 170.0 161.2 160.330 142.170 126.670 16.460 32.825<br />

Cercopithecus mitis 0.832 0.080 179.0 170.0 166.170 142.830 128.830 20.226 31.501<br />

Cheirogaleus major 0.664 181.0 186.8 176.750 154.000 108.750 13.818<br />

Chiropotes satanas 0.850 0.085 170.0 169.5 169.830 157.833 128.000 18.543 26.481<br />

Daubentonia madgascariensis 0.930 157.0 169.2 169.600 147.600 110.600 19.762<br />

Erythrocebus patas 0.824 0.080 167.0 158.7 152.170 134.833 125.500 22.377 35.374<br />

Eulemur fulvus 0.768 0.064 171.0 183.7 177.000 158.000 126.170 19.351 25.150<br />

Euoticus elegantulus 0.827 0.065 158.0 177.7 160.500 140.830 120.500 8.989 18.417<br />

Gorilla gorilla 1.021 0.086 148.0 157.3 163.000 150.200 124.670 43.777 48.110<br />

Hapalemur griseus 0.795 0.064 175.0 179.5 173.170 144.170 110.600 13.809 23.179<br />

Homo sapiens 1.635 0.118 112.0 137.5 163.330 153.500 123.670 42.751 50.598<br />

Hylobates lar 0.918 0.084 153.0 168.3 165.500 158.333 129.334 22.265 32.856<br />

Hylobates moloch 0.943 0.085 160.0 172.5 175.670 159.330 117.000 21.749 30.632<br />

Indri indri 0.704 0.060 168.0 186.2 175.670 166.330 124.200 25.183 29.202<br />

Lagothrix lagotricha 0.879 0.082 176.0 176.0 162.330 151.400 123.400 24.909 31.039<br />

Leontopithecus rosalia 0.804 0.076 173.0 172.5 177.500 154.500 125.500 11.211 19.259<br />

Lepilemur mustelinus 0.811 178.0 181.2 158.330 143.170 128.170 12.149<br />

Lophocebus albigena 0.877 0.085 163.0 161.5 187.000 163.750 144.000 24.181 30.611<br />

Loris tardigradus 0.848 0.079 162.0 187.0 152.500 119.667 117.167 10.432 13.090<br />

Macaca nigra 0.880 0.081 158.0 151.0 154.000 134.500 122.250 22.108 32.258<br />

Macaca sylvana 0.865 0.083 154.0 159.2 150.000 23.086 32.901<br />

M<strong>and</strong>rillus leucopaeus 0.842 0.079 160.0 155.0 154.600 130.600 123.000 33.170 38.816<br />

M<strong>and</strong>rillus sphinx 0.828 0.080 154.0 152.5 154.583 116.750 27.404 39.082<br />

Microcebus murinus 0.820 157.0 157.830 142.833 128.000<br />

Miopithecus talapoin 0.955 0.096 168.0 164.2 180.000 150.000 105.000 13.612 22.387<br />

Mirza coquereli 0.666 0.057 158.0 189.0 156.670 138.167 121.167 11.991 18.179<br />

Nasalis larvatus 0.890 158.0 188.170 161.330 129.330<br />

Nycticebus coucang 0.747 0.065 171.0 185.5 166.830 146.330 118.670 12.933 21.012<br />

(Continued)


Species or specimen IRE1 1 IRE5 2 (degrees)<br />

CBA4 1<br />

TABLE 7. (Continued)<br />

CBA1<br />

(degrees)<br />

AOA 1<br />

(degrees)<br />

AFK 1<br />

(degrees)<br />

Op-Ba clivus<br />

(degrees) 2<br />

Basion-sella<br />

(mm) 2<br />

Sella-foramen<br />

cecum (mm) 2<br />

Otolemur crassicaudatus 0.753 0.060 167.0 178.7 160.670 155.830 132.670 14.908 23.434<br />

Pan troglodytes 1.039 0.091 152.0 158.7 151.670 123.000 121.830 35.633 45.114<br />

Papio anubis 0.865 0.084 154.0 153.2 196.830 175.330 124.000 29.010 36.697<br />

Perodicticus potto 0.731 181.0 157.500 146.333 122.834 17.820<br />

Piliocolobus badius 0.891 0.080 156.0 166.5 176.670 161.000 115.500 21.070 30.809<br />

Pithecia pithecia 0.771 0.074 182.0 175.8 167.400 162.000 116.000 17.670 25.883<br />

Pongo pygmaeus 1.026 0.085 135.0 160.8 147.000 138.830 115.000 41.228 44.834<br />

Semnopithecus entellus 0.906 0.081 148.0 159.8 150.830 145.330 122.830 21.870 33.094<br />

Presbytis melalophos 0.917 0.087 154.0 160.7 151.830 143.500 120.670 18.318 28.798<br />

Presbytis rubicunda 0.886 0.087 152.0 159.0 159.200 143.750 124.750 20.201 28.047<br />

Procolobus verus 0.874 0.081 157.0 171.4 166.830 154.500 121.670 18.631 28.264<br />

Propithecus verreauxi 0.788 0.068 168.0 184.8 148.570 137.000 117.714 20.548 25.056<br />

Pygathrix nemaeus 0.933 0.087 145.0 156.1 151.330 143.500 120.833 20.451 30.254<br />

Rhinopithecus roxellanae 1.002 0.089 150.0 157.2 154.500 143.830 121.340 21.601 32.294<br />

Saguinus fuscicollis 0.769 0.075 170.0 164.5 155.670 140.000 113.500 10.395 17.039<br />

Saimiri sciureus 0.926 0.087 169.0 165.5 153.170 135.330 118.000 12.046 20.672<br />

Simias concolor 0.786 0.079 161.0 160.2 174.670 160.833 129.167 19.658 28.526<br />

Symphalangus syndactylus 0.868 0.078 167.0 173.7 139.000 121.750 115.000 27.721 37.234<br />

Tarsius syrichta 1.005 0.062 141.0 151.0 149.000 133.000 119.833 6.099 14.796<br />

<strong>The</strong>ropithecus gelada 0.841 0.084 152.0 156.5 163.000 146.833 123.000 26.796 33.425<br />

Trachypithecus cristata 0.857 0.076 166.0 168.0 176.170 157.500 117.670 22.247 30.804<br />

Varecia variegata 0.752 0.062 177.0 187.8 21.726 29.495<br />

WT 17000 0.088 156.0<br />

OH5 0.098 135.0<br />

Sts 5 1.292 0.100 112.5 147.0 36.700 41.700<br />

1 Data primarily from Ross <strong>and</strong> Ravosa (1993).<br />

2 New data collected for this study from same radiographic collection as for Ross <strong>and</strong> Ravosa (1993).


148 YEARBOOK OF PHYSICAL ANTHROPOLOGY [Vol. 43, 2000<br />

tion, <strong>and</strong> also to study potential interactions<br />

between <strong>cranial</strong> <strong>base</strong> shape <strong>and</strong> facial shape<br />

in <strong>primate</strong>s <strong>and</strong> other mammals.<br />

THE CRANIAL BASE AND POSTURE<br />

Apart from a few recent studies reviewed<br />

here, little is known about the relationship<br />

between basi<strong>cranial</strong> morphology, head <strong>and</strong><br />

neck posture, <strong>and</strong> other aspects of head<br />

morphology related to locomotion. One persistent<br />

hypothesis that is especially relevant<br />

to hominin evolution is that flexion of<br />

the <strong>cranial</strong> <strong>base</strong> is an adaptation for orthograde<br />

posture in hominins because it causes<br />

the foramen magnum to have a relatively<br />

more anterior position <strong>and</strong> ventral orientation<br />

(Bolk, 1909, 1910; Duckworth, 1915;<br />

Weidenreich, 1941; Schultz, 1942, 1955;<br />

Ashton <strong>and</strong> Zuckerman, 1952, 1956; Ashton,<br />

1952; Moore et al., 1973; Adams <strong>and</strong><br />

Moore, 1975; DuBrul, 1977, 1979; Dean <strong>and</strong><br />

Wood, 1981, 1982). Indeed, these features<br />

are often invoked in attempts to reconstruct<br />

head posture in fossil hominins (e.g., White<br />

et al., 1994). All <strong>primate</strong>s have the center of<br />

mass of the head located anterior to the<br />

occipital condyles, such that more anteriorly<br />

positioned occipital condyles relative to<br />

head length reduce the lever arm between<br />

the center of mass <strong>and</strong> the atlanto-occipital<br />

joint. This balancing, in turn, reduces torque<br />

about this joint, thereby reducing the magnitude<br />

of the force required from the nuchal<br />

muscles to hold up the head (Schultz, 1942).<br />

<strong>The</strong> occipital condyles can be moved rostrally<br />

relative to overall head length by flexing<br />

the basicranium <strong>and</strong>/or shortening the<br />

posterior <strong>cranial</strong> <strong>base</strong>. In vitro experiments<br />

by Demes (1985) demonstrated that the<br />

more ventral orientation of rostrally placed<br />

occipital condyles orients the articular surfaces<br />

closer to perpendicular to the compressive<br />

force acting through the center of mass<br />

of the head, potentially reducing shearing<br />

forces acting across the joint that need to be<br />

resisted by muscles or ligaments.<br />

<strong>The</strong>re are some experimental data that<br />

rodents forced to walk bipedally (Moss,<br />

1961; Fenart, 1966; Riesenfeld, 1966) develop<br />

more flexed <strong>cranial</strong> <strong>base</strong>s. However,<br />

the hypothesis that variations in <strong>cranial</strong><br />

<strong>base</strong> angle are adaptations for head posture,<br />

however, is not well supported by comparative<br />

data. Among <strong>primate</strong>s, basi<strong>cranial</strong> flexion<br />

has been shown to be uncorrelated with<br />

either qualitative estimates of body posture<br />

(Ross <strong>and</strong> Ravosa, 1993) or quantitative<br />

measures of head <strong>and</strong> neck orientation<br />

(Strait <strong>and</strong> Ross, 1999). <strong>The</strong> partial correlation<br />

analysis of Strait <strong>and</strong> Ross (1999) confirmed<br />

relative brain size as a more important<br />

determinant of variation in basi<strong>cranial</strong><br />

angle, even when facial orientation <strong>and</strong><br />

head <strong>and</strong> neck posture were taken into account.<br />

Although foramen magnum orientation<br />

relative to anterior <strong>cranial</strong> <strong>base</strong> orientation<br />

(S-FC) has been shown to be related<br />

to relative brain size (Biegert, 1963; Spoor,<br />

1997), the relationship between head <strong>and</strong><br />

neck posture <strong>and</strong> foramen magnum orientation<br />

has not yet been evaluated.<br />

We gathered data on foramen magnum<br />

orientation (FM) from the same radiographs<br />

used by Ross <strong>and</strong> Ravosa (1993) <strong>and</strong> Ravosa<br />

(1991b), <strong>and</strong> combined these data with measures<br />

of hominids reported by Spoor (1993)<br />

<strong>and</strong> with measures of head <strong>and</strong> neck orientation<br />

reported by Strait <strong>and</strong> Ross (1999)<br />

(Fig. 13). <strong>The</strong> head-neck angle is the angle<br />

between neck inclination <strong>and</strong> orbit inclination,<br />

both relative to the substrate (Strait<br />

<strong>and</strong> Ross, 1999). <strong>The</strong> values for FM orientation<br />

relative to the clivus (FM CO) are in<br />

Table 7 <strong>and</strong> are summarized in Figure 14.<br />

<strong>The</strong> values for FM orientation relative to<br />

the orbital axis were calculated from measures<br />

of the orientation of both these planes<br />

relative to the clivus ossis occipitalis (CO) as<br />

180°-AOA-FM CO. <strong>The</strong>se data show that<br />

FM CO is not significantly correlated<br />

with basi<strong>cranial</strong> flexion, orbital axis orientation,<br />

the orientation of the head relative<br />

to the neck, or the size of the cerebellum<br />

relative to the posterior basicranium (Table<br />

8). Nor is foramen magnum orientation relative<br />

to the orbital axis (FM AOA) correlated<br />

with any of these variables, except<br />

AOA (Table 8). Of particular interest is the<br />

lack of correlation between the head-neck<br />

angle of Strait <strong>and</strong> Ross (1999) (Fig. 14),<br />

suggesting that foramen magnum orientation<br />

is not a good indicator of the orientation<br />

of the neck during habitual locomotion. This<br />

calls into question attempts to estimate<br />

head <strong>and</strong> neck posture from data on foramen<br />

magnum orientation in fossils.


D.E. Lieberman et al.]<br />

PRIMATE CRANIAL BASE 149<br />

Fig. 14. Bivariate plot of foramen magnum orientation relative to orbital axis (new data) against<br />

head-neck angle (Strait <strong>and</strong> Ross, 1999). Homo has unusually airorhynch (dorsally rotated) orbits<br />

relative to its foramen magnum orientation. Across <strong>primate</strong>s the relationship between these variables is<br />

significant (r 0.541, P 0.05); across nonhuman <strong>primate</strong>s r 0.691 (P 0.01).<br />

TABLE 8. Pearson’s r <strong>and</strong> P-values for correlations<br />

between measures of foramen magnum orientation <strong>and</strong><br />

measures of head morphology <strong>and</strong> orientation<br />

FM AOA FM CO<br />

r P r P<br />

CBA1 0.482 0.2 0.160 0.6<br />

CBA4 0.145 0.7 0.258 0.5<br />

AOA 0.841 0.001* 0.378 0.3<br />

Head-neck angle 0.405 0.2 0.202 0.6<br />

Cerebellum/B-a-se 0.169 0.6 0.170 0.6<br />

* P 0.05.<br />

Another widely accepted notion holds that<br />

a shift to orthogrady necessitates ventral<br />

flexion of the face, particularly the orbits,<br />

relative to the rest of the skull so that animals<br />

can continue to look rostrally (reviewed<br />

in Ross, 1995a; Strait <strong>and</strong> Ross,<br />

1999). Strait <strong>and</strong> Ross (1999) found that<br />

most <strong>primate</strong>s orient their orbits slightly<br />

downwards during locomotion. When other<br />

variables are controlled for using partial<br />

correlation analysis, orbital axis orientation<br />

relative to the posterior <strong>cranial</strong> <strong>base</strong> correlates<br />

significantly with head <strong>and</strong> neck posture.<br />

<strong>The</strong>se findings suggest that the direction<br />

of gaze cannot be reoriented through<br />

evolution solely by alterations in neck orientation<br />

or head orientation relative to the<br />

neck, <strong>and</strong> that there is a relationship between<br />

locomotor behavior <strong>and</strong> orbit orientation<br />

relative to the rest of the skull.<br />

This latter conclusion also explains why<br />

the orientation of the lateral semicircular<br />

canal (LSC) may correlate with head posture<br />

(Fenart <strong>and</strong> Pellerin, 1988 <strong>and</strong> references<br />

therein). Humans <strong>and</strong> animals of<br />

varying postures habitually hold their<br />

heads with the LSCs pitched upwards by<br />

several degrees (Sakka et al., 1976; Vidal et<br />

al., 1986; Graf et al., 1995). <strong>The</strong> semicircular<br />

canals <strong>function</strong> as accelerometers, registering<br />

changes in head velocity (or changes<br />

therein, i.e., acceleration), not head orientation<br />

relative to gravity (which is accomplished<br />

using the otolith organs). Thus, the<br />

intrinsic <strong>function</strong> of the canals does not explain<br />

any relationship between LSC orientation<br />

<strong>and</strong> head posture. An alternative,


150 YEARBOOK OF PHYSICAL ANTHROPOLOGY [Vol. 43, 2000<br />

Fig. 15. Bivariate plot of orbital axis orientation relative to clivus (AOA) (Ross <strong>and</strong> Henneberg, 1995)<br />

against lateral semicircular canal (LSC) orientation relative to a line from basion to nasion (Spoor, 1993).<br />

<strong>The</strong> two variables are significantly correlated across all <strong>primate</strong>s (r 0.745, P 0.01) <strong>and</strong> nonhuman<br />

<strong>primate</strong>s (r 0.906, P 0.0001). Homo has unusually airorhynch (dorsally rotated) orbits for its LSC<br />

orientation (see also Ross <strong>and</strong> Henneberg, 1995; Strait <strong>and</strong> Ross, 1999).<br />

more likely answer lies in the role of the<br />

semicircular canals in vestibulo-ocular reflexes<br />

(VORs). VORs maintain a stable retinal<br />

image during head movements by coordinating<br />

eye movements with changes in<br />

head velocity. To achieve this, each semicircular<br />

canal is wired up to two extraocular<br />

muscles via a three-neuron reflex arc. 6 Because<br />

reflex arcs are evolutionarily conservative<br />

(Graf, 1988), each semicircular canal<br />

must remain roughly parallel with the line<br />

of action of the two extraocular muscles to<br />

which it provides primary excitatory input<br />

during VORs. This <strong>function</strong>al constraint<br />

predicts that evolutionary changes in medial<br />

<strong>and</strong> lateral rectus orientation will be<br />

associated with changes in LSC orientation.<br />

Because the medial <strong>and</strong> lateral recti arise<br />

6 Each posterior semicircular canal is connected to its ipsilateral<br />

superior oblique <strong>and</strong> contralateral inferior rectus, each anterior<br />

semicircular canal to its ipsilateral superior rectus <strong>and</strong> its<br />

contralateral inferior oblique, <strong>and</strong> each LSC to its ipsilateral<br />

medial rectus <strong>and</strong> contralateral lateral rectus (K<strong>and</strong>el et al.,<br />

1991).<br />

from the annulus tendineus around the optic<br />

canal <strong>and</strong> insert on the equator of the<br />

eyeball, they lie roughly in the same transverse<br />

plane as the orbital axis (see Measurement<br />

Definitions). Thus, LSC orientation<br />

should be correlated with orbital axis orientation.<br />

As an initial test of this hypothesis, Figure<br />

15 plots AOA (orbital axis orientation<br />

relative to CO) (data from Ross <strong>and</strong> Henneberg,<br />

1995) against LSC orientation relative<br />

to a line from basion to sella (ba-s)<br />

(Spoor <strong>and</strong> Zonneveld, 1998). Because CO is<br />

very similar to the line from ba-s in most<br />

taxa, these two angles roughly reflect the<br />

orientations of the LSC <strong>and</strong> the orbital axis<br />

both relative to the occipital clivus. Although<br />

these data derive from different<br />

specimens, a significant relationship is observed:<br />

LSC orientation is closely correlated<br />

with orbit orientation (r 0.824; the negative<br />

correlation reflects the fact that the<br />

LSC ba-s angle is measured above the<br />

LSC line, while the AOA angle is measured


D.E. Lieberman et al.]<br />

PRIMATE CRANIAL BASE 151<br />

below the orbital axis line). <strong>The</strong>se data<br />

therefore lend support to the hypothesis<br />

that animals orient their heads at rest (<strong>and</strong><br />

during locomotion) primarily to orient their<br />

eyes towards the horizon rather than to orient<br />

their semicircular canals in any particular<br />

way. In sum, orbital axis orientation<br />

relative to the clivus <strong>and</strong> foramen magnum<br />

orientation relative to the orbital axis are<br />

hypothesized to be evolutionary adaptations<br />

to head posture during habitual locomotion,<br />

<strong>and</strong> appear necessary because of the limited<br />

range of motion at the atlanto-occipital joint<br />

(Graf et al., 1995).<br />

MAJOR UNRESOLVED ISSUES OF<br />

CRANIAL BASE VARIATION IN PRIMATE<br />

EVOLUTION<br />

Studies of <strong>cranial</strong> <strong>base</strong> variation in fossil<br />

<strong>primate</strong>s <strong>and</strong> hominins have been rare because<br />

this region of the skull is usually<br />

poorly preserved or destroyed in most fossils,<br />

<strong>and</strong> because it is hard to visualize or<br />

measure without radiographs or computed<br />

tomography (CT) scans. <strong>The</strong> coming years,<br />

however, are likely to see a renaissance of<br />

research on the role of the <strong>cranial</strong> <strong>base</strong> in<br />

<strong>primate</strong> <strong>cranial</strong> evolution as CT scans of<br />

fossils become more readily available. Here<br />

we review six major topics where future research<br />

on the <strong>cranial</strong> <strong>base</strong> in both fossil <strong>and</strong><br />

extant <strong>primate</strong>s promises to provide important<br />

insights: 1) the relationship between<br />

encephalization, circumorbital form, <strong>and</strong><br />

the origin of <strong>primate</strong>s; 2) the evolution of<br />

the integrated “facial block” in haplorhines;<br />

3) the determinants of basi<strong>cranial</strong> flexion in<br />

hominins; 4) the relationship between basi<strong>cranial</strong><br />

flexion <strong>and</strong> the shape of the vocal<br />

tract; 5) the role of the <strong>cranial</strong> <strong>base</strong> in facial<br />

retraction in Homo sapiens; <strong>and</strong> 6) the reliability<br />

of basi<strong>cranial</strong> characters as indicators<br />

of <strong>primate</strong> phylogeny.<br />

Primate origins, encephalization, <strong>and</strong><br />

circumorbital form<br />

<strong>The</strong> basicranium likely played a key role<br />

in the evolution of the unique configuration<br />

of the <strong>primate</strong> skull. Over the past three<br />

decades, the visual predation hypothesis<br />

(VPH) has become a well-accepted model of<br />

<strong>primate</strong> origins (Martin, 1990; Fleagle,<br />

1999). <strong>The</strong> VPH argues that the first <strong>primate</strong>s<br />

were nocturnal visual predators of<br />

small invertebrates <strong>and</strong> vertebrates, <strong>and</strong><br />

this required more anteriorly facing <strong>and</strong><br />

closely approximated orbital apertures<br />

(Cartmill, 1970, 1972, 1974, 1992). Increased<br />

orbital convergence enlarges the<br />

binocular field for greater stereoscopic vision<br />

<strong>and</strong> a clear retinal image for depth<br />

perception <strong>and</strong> prey location (Allman, 1977,<br />

1982). <strong>The</strong> VPH further posits that relatively<br />

larger orbits <strong>and</strong> grasping appendages<br />

with nails are adaptations to being<br />

nocturnal in an arboreal, terminal-branch<br />

setting (Cartmill, 1970, 1972, 1974, 1992;<br />

Kay <strong>and</strong> Cartmill, 1974, 1977; Dagosto,<br />

1988; Covert <strong>and</strong> Hamrick, 1993; Hamrick,<br />

1998, 1999; Lemelin, 1999). <strong>The</strong>se adaptations<br />

differ significantly from the <strong>cranial</strong><br />

<strong>and</strong> locomotor specializations of putative<br />

sister taxa such as plesiadapiforms <strong>and</strong> dermopterans<br />

(Cartmill, 1972, 1974, 1992; Kay<br />

<strong>and</strong> Cartmill, 1974, 1977; Beard, 1993; Ravosa<br />

et al., 2000a).<br />

According to the VPH, increased orbital<br />

convergence moves the orbital apertures out<br />

of the plane of the temporal fossa, a condition<br />

entailing greater ocular disruption during<br />

mastication (Cartmill, 1970, 1972, 1974,<br />

1992). A rigid postorbital bar may <strong>function</strong><br />

to stiffen the lateral orbital margins <strong>and</strong><br />

thus counter ocular deformation during biting<br />

<strong>and</strong> chewing to ensure a high level of<br />

stereoscopic acuity in an organism that processes<br />

food while hunting <strong>and</strong> foraging<br />

(Cartmill, 1970, 1972). This appears to be<br />

particularly important, given that strepsirhines<br />

with unfused symphyses have been<br />

shown to recruit relatively less balancingside<br />

than working-side adductor muscle<br />

force during unilateral mastication (Hyl<strong>and</strong>er<br />

et al., 1998, 2000). This differential<br />

muscle recruitment results in a pattern of<br />

lower strains along the balancing-side postorbital<br />

bar than the working-side postorbital<br />

bar (Ravosa et al., 2000a). Thus, an<br />

organism with a postorbital ligament <strong>and</strong> a<br />

stepsirhine-like adductor pattern (the latter<br />

of which is inferred for basal <strong>primate</strong>s <strong>base</strong>d<br />

on the presence of unfused symphyses; Ravosa,<br />

1996, 1999) would experience an<br />

asymmetrical circumorbital <strong>and</strong>, in turn, an<br />

ocular loading pattern that is hypothesized


152 YEARBOOK OF PHYSICAL ANTHROPOLOGY [Vol. 43, 2000<br />

to compromise effective stereoscopic visual<br />

acuity.<br />

Analyses of felids (perhaps the best analog<br />

for the skull of basal <strong>primate</strong>s), herpestids,<br />

<strong>and</strong> pteropodids demonstrate that<br />

postorbital bar formation characterizes taxa<br />

with increased orbital convergence <strong>and</strong>/or<br />

greater orbital frontation (Noble et al.,<br />

2000; Ravosa et al., 2000a). As both orbital<br />

characteristics became more developed during<br />

early <strong>primate</strong> evolution (Simons, 1962;<br />

Cartmill, 1970, 1972, 1974, 1992; Szalay et<br />

al., 1987; Martin, 1990; Fleagle, 1999), it is<br />

likely that <strong>primate</strong> postorbital bar development<br />

is correlated with both evolutionary<br />

transformations in the orbital complex (Ravosa<br />

et al., 2000a).<br />

<strong>The</strong> role of orbital frontation in postorbital<br />

bar formation is especially significant<br />

because it apparently reflects interactions<br />

among several factors unique to the early<br />

evolution of small <strong>primate</strong>s (Ravosa et al.,<br />

2000a). First among these factors is encephalization.<br />

Both early <strong>primate</strong>s <strong>and</strong> their putative<br />

ancestors were tiny, weighing between<br />

100–300 g (Kay <strong>and</strong> Cartmill, 1977;<br />

Dagosto, 1988; Martin, 1990; Beard, 1993;<br />

Covert <strong>and</strong> Hamrick, 1993; Fleagle, 1999).<br />

Thus, the greater encephalization of basal<br />

<strong>primate</strong>s (Martin, 1990) did not result from<br />

phyletic size decreases 7 (Gould, 1975; Shea,<br />

1987). Instead, increased relative brain size<br />

<strong>and</strong> orbital frontation appear linked to nocturnal<br />

visual predation (Ravosa et al.,<br />

2000a), which is related to increases in the<br />

relative size of the visual cortex (Cartmill,<br />

1974, 1992) <strong>and</strong> greater encephalization<br />

(Barton, 1998). This hypothesis complements<br />

prior claims that increased encephalization<br />

in basal <strong>primate</strong>s is related to the<br />

unique combination of arboreality, precociality,<br />

<strong>and</strong> small body size (Shea, 1987).<br />

Apart from being more encephalized than<br />

non<strong>primate</strong> archontans, the first <strong>primate</strong>s<br />

(<strong>and</strong> felids) possessed relatively larger eyes<br />

<strong>and</strong> more convergent orbits (Kay <strong>and</strong> Cartmill,<br />

1977; Martin, 1990; Covert <strong>and</strong> Hamrick,<br />

1993; Cartmill, 1992; Ravosa et al.,<br />

7 As phyletic dwarfs are more encephalized than like-sized<br />

sister taxa (Gould, 1975; Shea, 1987), presumably they are also<br />

more flexed, highlighting the complex relationships between<br />

brain evolution <strong>and</strong> changes in basi<strong>cranial</strong> <strong>and</strong> facial form.<br />

2000a). This combination of derived features<br />

creates a spatial packing problem in<br />

which the position of the frontal lobes <strong>and</strong><br />

anterior <strong>cranial</strong> fossa have a major influence<br />

on orbital aperture orientation, integrating<br />

morphological variation in these adjacent<br />

<strong>cranial</strong> regions (see above). Because<br />

neural <strong>and</strong> ocular size scale with negative<br />

allometry (Schultz, 1940; Gould, 1975), this<br />

structural constraint would be particularly<br />

marked in smaller, <strong>and</strong> thus more frontated,<br />

taxa—the morphospace in which<br />

basal <strong>primate</strong>s <strong>and</strong> felids happen to develop<br />

postorbital bars.<br />

Anthropoid origins <strong>and</strong> <strong>cranial</strong><br />

<strong>base</strong>-face interactions<br />

Another interesting problem is the role of<br />

the <strong>cranial</strong> <strong>base</strong> in the fundamental rearrangement<br />

of the face along the stem lineage<br />

leading to haplorhines <strong>and</strong> anthropoids.<br />

<strong>The</strong>se changes include: increased<br />

orbital convergence <strong>and</strong> frontation; increased<br />

relative brain size, especially in the<br />

frontal <strong>and</strong> temporal lobes; a more acutely<br />

angled (flexed) <strong>cranial</strong> <strong>base</strong>; partial retraction<br />

of the face below the braincase; reduction<br />

of the olfactory apparatus <strong>and</strong><br />

interorbital region with concomitant<br />

approximation of the orbits; relatively<br />

smaller orbits associated with diurnality;<br />

<strong>and</strong> the presence of a postorbital septum<br />

(Cartmill, 1970, 1980; Radinsky, 1968,<br />

1979; Rosenberger, 1985, 1986; Ravosa,<br />

1991b; Ross <strong>and</strong> Ravosa, 1993; Ross,<br />

1995a,b). Not surprisingly, changes in <strong>cranial</strong><br />

<strong>base</strong> shape have been key features of all<br />

the models proposed to explain the order in<br />

which these features arose, their putative<br />

<strong>function</strong>s, <strong>and</strong> their interactions. According<br />

to Cartmill (1970), the fundamental difference<br />

between haplorhines <strong>and</strong> strepsirhines<br />

centers on whether orbital convergence occurs<br />

towards the top of the skull or at the<br />

front. If, as in haplorhines, the orbits converge<br />

anteriorly (associated with increased<br />

orbital frontation), then the orbits not only<br />

compress the olfactory region, reducing its<br />

dimensions, but also become closely approximated<br />

to the midline anterior <strong>cranial</strong> <strong>base</strong>.<br />

This necessarily reduces the size of the olfactory<br />

apparatus <strong>and</strong> forms an interorbital<br />

septum, facilitating a subsequent increase


D.E. Lieberman et al.]<br />

PRIMATE CRANIAL BASE 153<br />

in relative encephalization by expansion of<br />

the brain forward over the top of the orbits.<br />

It has been argued that, as predicted by<br />

Dabelow (1931), by bringing the orbits <strong>and</strong><br />

anterior <strong>cranial</strong> <strong>base</strong> into close continuity,<br />

they became structurally integrated such<br />

that changes in the orientation of one necessarily<br />

affect the other (Ross <strong>and</strong> Ravosa,<br />

1993). Small nocturnal strepsirhines, in<br />

contrast, tend to converge their orbits towards<br />

the top of their head, not compressing<br />

the olfactory apparatus, <strong>and</strong> preventing<br />

brain expansion over the orbits concomitant<br />

with a shift to diurnality. <strong>The</strong> orbits <strong>and</strong><br />

anterior <strong>cranial</strong> <strong>base</strong> in all strepsirhines are<br />

not as unified structurally, <strong>and</strong> thus basi<strong>cranial</strong><br />

flexion <strong>and</strong> orbit orientation do not<br />

covary as greatly.<br />

<strong>The</strong> evidence that all omomyiforms that<br />

are well enough preserved have an interorbital<br />

septum below the olfactory tract, like<br />

that of Tarsius <strong>and</strong> small anthropoids (Ross,<br />

1994), suggests that this integration of anterior<br />

<strong>cranial</strong> <strong>base</strong> <strong>and</strong> upper face may be an<br />

early unique derived feature of the haplorhine<br />

stem lineage. This integration set the<br />

stage for the subsequent evolution of the<br />

haplorhine postorbital septum. Increased<br />

orbital frontation (klinorhynchy) <strong>and</strong> convergence<br />

in early anthropoids caused the<br />

anterior temporal muscles <strong>and</strong> orbit to come<br />

into close proximity, leading to the evolution<br />

of a postorbital septum (Ross, 1995b, 1996).<br />

Exactly why increased frontation occurred<br />

in haplorhines remains unexplained, although<br />

it has been argued that increased<br />

relative brain size might have caused increased<br />

basi<strong>cranial</strong> flexion, producing increased<br />

orbital frontation as a consequence<br />

of the integration of the anterior <strong>cranial</strong><br />

<strong>base</strong> <strong>and</strong> orbits (Ross, 1996; see also Ravosa<br />

et al., 2000a on the effects of small size on<br />

skull form in basal <strong>primate</strong>s). This hypothesis<br />

remains to be evaluated in fossil stem<br />

anthropoids for want of well-preserved specimens,<br />

but is not supported by the fact that<br />

early anthropoids had a postorbital septum<br />

in conjunction with relatively small brains<br />

(Ross, 2000). It remains to be determined<br />

why the haplorhine stem lineage evolved<br />

more frontated orbits, but the findings of<br />

Strait <strong>and</strong> Ross (1999) suggest that the role<br />

of posture should be considered.<br />

Variation in hominin <strong>cranial</strong> <strong>base</strong> angle<br />

How to account for variation in basi<strong>cranial</strong><br />

flexion in hominins has been controversial<br />

<strong>and</strong> remains unresolved. Several recent<br />

studies (e.g., Ross <strong>and</strong> Henneberg, 1995;<br />

Spoor, 1997) have shown that the <strong>cranial</strong><br />

<strong>base</strong> in Australopithecus <strong>and</strong> Homo is generally<br />

more flexed than in Pan or other nonhuman<br />

<strong>primate</strong>s. This difference raises two<br />

questions. First, how much is <strong>cranial</strong> <strong>base</strong><br />

flexion among early hominins related to upright<br />

posture, facial orientation, or brain<br />

size? Second, what factors account for the<br />

observed variation in <strong>cranial</strong> <strong>base</strong> angle<br />

among hominins?<br />

Although <strong>cranial</strong> <strong>base</strong> angles vary substantially<br />

among hominin species, australopithecines<br />

have CBAs generally intermediate<br />

between humans <strong>and</strong> chimpanzees, but<br />

with more flexed <strong>cranial</strong> <strong>base</strong>s among A.<br />

boisei <strong>and</strong> less flexed <strong>cranial</strong> <strong>base</strong>s for A.<br />

aethiopicus (KNM-WT 17000) (F. Spoor,<br />

personal communication). Until now, the<br />

hypothesis that <strong>cranial</strong> <strong>base</strong> flexion in hominins<br />

is an adaptation for increased brain<br />

size relative to basi<strong>cranial</strong> length (Gould,<br />

1977) has received the most support (Ross<br />

<strong>and</strong> Ravosa, 1993; Ross <strong>and</strong> Henneberg,<br />

1995; Spoor, 1997; McCarthy, 2001). In particular,<br />

if one measures IRE5 vs. CBA1<br />

(thereby including the cribriform plate in<br />

measures of both basi<strong>cranial</strong> length <strong>and</strong> the<br />

angle of the <strong>cranial</strong> <strong>base</strong>), then H. sapiens<br />

<strong>and</strong> other hominin taxa such as A. africanus<br />

have exactly the degree of flexion expected<br />

by basi<strong>cranial</strong> length.<br />

However, not all of the variation in hominin<br />

CBA can be explained by relative brain<br />

size. For example, Ne<strong>and</strong>erthals <strong>and</strong> archaic<br />

Homo fossils such as Kabwe have considerably<br />

more extended CBA1s than H. sapiens<br />

(about 15°), even though they are<br />

bipedal <strong>and</strong> similarly encephalized (Ruff et<br />

al., 1997). In addition, other measures of<br />

<strong>cranial</strong> <strong>base</strong> angle <strong>and</strong> relative encephalization,<br />

which do not include the cribriform<br />

plate (CBA4 <strong>and</strong> IRE5), indicate that H.<br />

sapiens have less flexed <strong>cranial</strong> <strong>base</strong>s then<br />

expected for anthropoids of their size (Ross<br />

<strong>and</strong> Henneberg, 1995). This finding highlights<br />

the likelihood that no single explanation<br />

will account for interspecific differences


154 YEARBOOK OF PHYSICAL ANTHROPOLOGY [Vol. 43, 2000<br />

in CBA among hominins. Other factors that<br />

have been argued to contribute to variation<br />

in CBA, including facial kyphosis <strong>and</strong> head<br />

<strong>and</strong> neck posture, cannot be shown to be<br />

important correlates of basi<strong>cranial</strong> flexion.<br />

Most importantly, the comparative study by<br />

Strait <strong>and</strong> Ross (1999) of CBA, head-neck<br />

angle, <strong>and</strong> orbital axis orientation among<br />

extant <strong>primate</strong>s found that basi<strong>cranial</strong> flexion<br />

is primarily influenced by relative brain<br />

size <strong>and</strong> not head <strong>and</strong> neck posture. <strong>The</strong><br />

traditional view that human basi<strong>cranial</strong><br />

flexion is somehow an adaptation for orthograde<br />

posture is no longer tenable.<br />

Cranial <strong>base</strong> flexion <strong>and</strong> vocal tract<br />

shape in hominins<br />

Flexion of the <strong>cranial</strong> <strong>base</strong> has been argued<br />

to be an important correlate of the<br />

shape of the vocal tract. In particular, several<br />

researchers (Lieberman et al., 1972,<br />

1992; Laitman <strong>and</strong> Crelin, 1976; Laitman et<br />

al., 1978, 1979; Laitman <strong>and</strong> Heimbuch,<br />

1982) claimed that the degree of flexion of<br />

the external <strong>cranial</strong> <strong>base</strong> (between the inferior<br />

aspect of the basioccipital clivus <strong>and</strong> the<br />

palate) influences the position of the hyolaryngeal<br />

complex. In most mammals, the<br />

hyoid <strong>and</strong> larynx lie close to the soft palate<br />

so that the epiglottis <strong>and</strong> soft palate can<br />

engage, permitting simultaneous breathing<br />

<strong>and</strong> swallowing (Laitman <strong>and</strong> Reidenberg,<br />

1993; German et al., 1996; Larson <strong>and</strong> Herring,<br />

1996; Crompton et al., 1997). In contrast,<br />

humans have a uniquely shaped pharynx<br />

in which the hyo-laryngeal complex<br />

descends inferiorly relative to the oral cavity,<br />

causing the soft palate <strong>and</strong> epiglottis to<br />

become disengaged so that the trachea <strong>and</strong><br />

esophagus share a common passageway<br />

(Negus, 1949; Crelin, 1973). Hyo-laryngeal<br />

descent is an important physiological basis<br />

for many aspects of human speech. Flexion<br />

of the external <strong>cranial</strong> <strong>base</strong> in combination<br />

with a low position of the larynx relative to<br />

the palate divide the human vocal tract (VT)<br />

into separate horizontal <strong>and</strong> vertical “tubes”<br />

of approximately equal length, whose crosssectional<br />

areas can be modified independently<br />

at least tenfold by the tongue<br />

(Stevens <strong>and</strong> House, 1955; Fant, 1960;<br />

Stevens, 1972; Baer et al., 1991). Dynamic<br />

filtering in the two-tube human VT <strong>function</strong>s<br />

to produce a wide range of vowels<br />

whose formant frequencies are acoustically<br />

distinct regardless of vocal tract length<br />

(Fant, 1960; Nearey, 1978; Lieberman,<br />

1984; Carre et al., 1994; Beckman et al.,<br />

1995).<br />

Although the speech-related acoustical<br />

properties of the unique human vocal tract<br />

are not in dispute, the role of <strong>cranial</strong> <strong>base</strong><br />

flexion in hyo-laryngeal descent is controversial,<br />

yet potentially important for reconstructing<br />

the anatomy of the vocal tract in<br />

fossil hominins Two questions are of special<br />

importance. First, what is the relationship<br />

of external <strong>cranial</strong> <strong>base</strong> flexion to the descent<br />

<strong>and</strong>/or positioning of the hyoid relative<br />

to the oral cavity? Second, how does the<br />

angle of the external <strong>cranial</strong> <strong>base</strong> change<br />

during development, <strong>and</strong> what is its relationship<br />

to internal <strong>cranial</strong> <strong>base</strong> angle?<br />

<strong>The</strong> original hypothesis that flexion of the<br />

external <strong>cranial</strong> <strong>base</strong> in humans contributes<br />

to laryngeal descent, <strong>and</strong> thus can be used<br />

to reconstruct the vocal tract of fossil hominins<br />

(Lieberman <strong>and</strong> Crelin, 1971; Lieberman<br />

et al., 1972; Laitman <strong>and</strong> Crelin, 1976;<br />

Laitman <strong>and</strong> Heimbuch, 1982; Lieberman,<br />

1984), was <strong>base</strong>d on three observations.<br />

First, these researchers inferred a relationship<br />

between <strong>cranial</strong> <strong>base</strong> flexion <strong>and</strong> laryngeal<br />

descent on the basis of comparisons of<br />

neonatal <strong>and</strong> adult humans vs. nonhuman<br />

<strong>primate</strong>s. In particular, human neonates,<br />

like nonhuman <strong>primate</strong> adults, have a relatively<br />

extended <strong>cranial</strong> <strong>base</strong> <strong>and</strong> a high hyoid,<br />

whereas only human adults have both a<br />

flexed <strong>cranial</strong> <strong>base</strong> <strong>and</strong> a descended larynx.<br />

Second, internal <strong>and</strong> external basi<strong>cranial</strong><br />

flexion were believed to covary with the descent<br />

of the larynx (George, 1978). Third,<br />

external <strong>cranial</strong> <strong>base</strong> flexion was believed to<br />

reorient the suprahyoid muscles <strong>and</strong> ligaments<br />

more vertically, <strong>and</strong> to shorten the<br />

antero-posterior length of the oropharynx<br />

between the palate <strong>and</strong> the cervical vertebrae,<br />

thus requiring the larynx <strong>and</strong> posterior<br />

tongue to be positioned lower relative to<br />

the hard palate.<br />

Lieberman <strong>and</strong> McCarthy (1999) tested<br />

the relationship between <strong>cranial</strong> <strong>base</strong> flexion<br />

<strong>and</strong> hyo-laryngeal position using a longitudinal<br />

series of radiographs in humans<br />

(the Denver Growth Study), which pre-


D.E. Lieberman et al.]<br />

PRIMATE CRANIAL BASE 155<br />

Fig. 16. Longitudinal changes in the angle of the<br />

external <strong>cranial</strong> <strong>base</strong> (CBA5 from basion-sphenobasionhormion)<br />

<strong>and</strong> the height of the vocal tract (the distance<br />

from the vocal folds to the plane of the hard palate,<br />

perpendicular to the posterior pharyngeal wall). Data<br />

are from a longitudinal study of growth in 15 males <strong>and</strong><br />

13 females (for details, see Lieberman <strong>and</strong> McCarthy,<br />

1999). Note that the height of the vocal tract continues<br />

to grow throughout the somatic growth period, whereas<br />

the external <strong>cranial</strong> <strong>base</strong> angle ceases to change appreciably<br />

after approximately 3 postnatal years.<br />

served details of larynx <strong>and</strong> hyoid position<br />

in upright subjects x-rayed during quiet respiration.<br />

This study found no statistically<br />

significant relationship between either internal<br />

or external <strong>cranial</strong> <strong>base</strong> flexion <strong>and</strong><br />

hyo-laryngeal descent. As Figure 16 shows,<br />

the <strong>cranial</strong> <strong>base</strong> <strong>and</strong> the position of the larynx<br />

must be partially independent in humans<br />

because the <strong>cranial</strong> <strong>base</strong> flexes rapidly<br />

during the first few years after birth,<br />

whereas the larynx <strong>and</strong> hyoid descend gradually<br />

until the end of the adolescent growth<br />

spurt. Consequently, the flexion of the external<br />

<strong>cranial</strong> <strong>base</strong> (<strong>and</strong> presumably its effects<br />

on suprahyoid muscle orientation) cannot<br />

be used to infer vocal tract dimensions<br />

in fossil hominins (Lieberman et al., 1998).<br />

Similarly, Chan (1991) demonstrated that<br />

the correlation between styloid process orientation<br />

<strong>and</strong> laryngeal position is not strong<br />

enough to estimate vocal tract dimensions<br />

reliably.<br />

External flexion has been measured in<br />

several ways, all of which suggest that external<br />

<strong>and</strong> internal <strong>cranial</strong> <strong>base</strong> angles are<br />

correlated with each other, but differ in<br />

their pattern of growth in humans <strong>and</strong> nonhuman<br />

<strong>primate</strong>s. Laitman et al. (1976,<br />

1978, 1979, 1982) developed a composite,<br />

size-corrected measure of exo<strong>cranial</strong> flexion<br />

between the basioccipital <strong>and</strong> the palate,<br />

which they measured on cross-sectional<br />

samples of humans, apes, monkeys, <strong>and</strong><br />

several fossil hominins. More recently,<br />

Lieberman <strong>and</strong> McCarthy (1999) measured<br />

external <strong>cranial</strong> <strong>base</strong> flexion in a longitudinal<br />

sample of humans using two lines, one<br />

extending from basion to sphenobasion, <strong>and</strong><br />

the second from sphenobasion to hormion.<br />

May <strong>and</strong> Sheffer (1999) took essentially the<br />

same measurement on cross-sectional samples<br />

of humans, chimpanzees, gorillas, <strong>and</strong><br />

several fossil hominins. <strong>The</strong>se studies all<br />

agree that flexion of the internal <strong>cranial</strong>


156 YEARBOOK OF PHYSICAL ANTHROPOLOGY [Vol. 43, 2000<br />

<strong>base</strong> in humans occurs prior to the eruption<br />

of the first permanent molars <strong>and</strong> then remains<br />

stable, but that the external <strong>cranial</strong><br />

<strong>base</strong> extends gradually in all nonhuman <strong>primate</strong>s<br />

throughout the period of facial<br />

growth. 8 Moreover, the patterns of external<br />

<strong>and</strong> internal <strong>cranial</strong> <strong>base</strong> angulation mirror<br />

each other. <strong>The</strong> internal <strong>cranial</strong> <strong>base</strong> (measured<br />

using both Ba-S-FC <strong>and</strong> L<strong>and</strong>zert’s<br />

angle) flexes in humans rapidly prior to 2<br />

postnatal years <strong>and</strong> then remains stable,<br />

but extends in nonhuman <strong>primate</strong>s gradually<br />

throughout the period of facial growth<br />

(Fig. 7) (Lieberman <strong>and</strong> McCarthy, 1999).<br />

Cranial <strong>base</strong> shape <strong>and</strong> facial<br />

projection in Homo<br />

Although the effects of <strong>cranial</strong> <strong>base</strong> angulation<br />

on the angle of the face relative to the<br />

rest of the skull (facial kyphosis) have long<br />

been the subject of much research (see<br />

above), there has been recent interest in the<br />

role of the <strong>cranial</strong> <strong>base</strong> on facial projection.<br />

Facial projection (which is a more general<br />

term for neuro-orbital disjunction) is defined<br />

here as the extent to which the nonrostral<br />

portion of the face is positioned anteriorly<br />

relative to the foramen caecum, the<br />

most antero-inferior point on the <strong>cranial</strong><br />

<strong>base</strong> (note that facial projection <strong>and</strong> prognathism<br />

are different). Variation in facial<br />

projection, along with an underst<strong>and</strong>ing of<br />

their developmental <strong>base</strong>s, may be important<br />

for testing hypotheses about recent<br />

hominin evolution. In particular, Lieberman<br />

(1995, 1998, 2000) has argued that<br />

variation in facial projection accounts for<br />

many of the major differences in overall<br />

craniofacial form between H. sapiens <strong>and</strong><br />

other closely-related “archaic” Homo taxa,<br />

including the Ne<strong>and</strong>erthals. Whereas all<br />

nonextant hominins have projecting faces,<br />

“anatomically modern” H. sapiens is<br />

uniquely characterized by a retracted facial<br />

profile in which the majority of the face lies<br />

beneath the braincase (Weidenreich, 1941;<br />

Moss <strong>and</strong> Young, 1960; Vinyard, 1994;<br />

Lieberman, 1995, 1998; Vinyard <strong>and</strong> Smith,<br />

1997; May <strong>and</strong> Sheffer, 1999; Ravosa et al.,<br />

2000b). As a consequence, H. sapiens also<br />

has a more vertical frontal profile, less projecting<br />

browridges, a rounder overall <strong>cranial</strong><br />

shape, <strong>and</strong> a relatively shorter oropharyngeal<br />

space between the back of the hard<br />

palate <strong>and</strong> the foramen magnum—virtually<br />

all of the supposed autapomorphies of “anatomically<br />

modern” H. sapiens.<br />

What is the role of the <strong>cranial</strong> <strong>base</strong> in<br />

facial projection? Lieberman (1998, 2000)<br />

proposed that four independent factors account<br />

for variation in facial projection: 1)<br />

antero-posterior facial length, 2) anterior<br />

<strong>cranial</strong> <strong>base</strong> length, 3) <strong>cranial</strong> <strong>base</strong> angle,<br />

<strong>and</strong> 4) the antero-posterior length of the<br />

middle <strong>cranial</strong> fossa from sella to PM<br />

plane. 9 Each of these variables has a different<br />

growth pattern, but combine to influence<br />

the position of the face relative to the<br />

basicranium <strong>and</strong> neurocranium. For example,<br />

facial projection can occur through having<br />

a long face relative to a short anterior<br />

<strong>cranial</strong> fossa, a long middle <strong>cranial</strong> fossa<br />

relative to the length of the anterior <strong>cranial</strong><br />

fossa, <strong>and</strong>/or a more extended <strong>cranial</strong> <strong>base</strong>.<br />

Partial correlation analyses of cross-sectional<br />

samples of Homo sapiens <strong>and</strong> Pan<br />

troglodytes indicate that each contributes<br />

significantly to the <strong>ontogeny</strong> of facial projection<br />

in humans <strong>and</strong> apes when the associations<br />

between these variables <strong>and</strong> with<br />

overall <strong>cranial</strong> length <strong>and</strong> endo<strong>cranial</strong> volume<br />

as well as other <strong>cranial</strong> dimensions are<br />

held constant (Lieberman, 2000). In other<br />

words, chimpanzees <strong>and</strong> humans with relatively<br />

longer faces, shorter anterior <strong>cranial</strong><br />

<strong>base</strong>s, less flexed <strong>cranial</strong> <strong>base</strong>s, <strong>and</strong>/or<br />

longer middle <strong>cranial</strong> fossae tend to have<br />

relatively more projecting faces.<br />

In an analysis of radiographs of fossil<br />

hominins, Lieberman (1998) argued that<br />

the major cause for facial retraction <strong>and</strong> its<br />

resulting effects on modern human <strong>cranial</strong><br />

shape was a change in the <strong>cranial</strong> <strong>base</strong><br />

rather than the face itself. Specifically, middle<br />

<strong>cranial</strong> fossa length (termed ASL) was<br />

estimated to be approximately 25% shorter<br />

in anatomically modern humans, both re-<br />

8 May <strong>and</strong> Sheffer (1999) did not find evidence for postnatal<br />

extension of the <strong>cranial</strong> <strong>base</strong> in Pan, largely because of insufficient<br />

sample sizes that were divided into overly large ontogenetic<br />

stages.<br />

9 This dimension was originally termed anterior sphenoid<br />

length (ASL), but it is really a measure of the midline prechordal<br />

length of the middle <strong>cranial</strong> fossa (Lieberman, 2000).


D.E. Lieberman et al.]<br />

PRIMATE CRANIAL BASE 157<br />

Fig. 17. Thin-plate spline analysis of an Australian male (from Queensl<strong>and</strong>) relative to Kabwe 1<br />

(target). Eighteen l<strong>and</strong>marks from each skull were initially superimposed using a resistant-fit Procrustes<br />

analysis. <strong>The</strong> deformation grid shows that the archaic Homo fossil has a relatively more projecting <strong>and</strong><br />

taller face, a more extended <strong>cranial</strong> <strong>base</strong>, a relatively shorter middle <strong>cranial</strong> fossa, <strong>and</strong> a relatively longer<br />

pharyngeal space between the palate <strong>and</strong> the foramen magnum.<br />

cent <strong>and</strong> Pleistocene, than in Ne<strong>and</strong>erthals<br />

<strong>and</strong> other taxa of archaic Homo, whereas<br />

anterior <strong>cranial</strong> <strong>base</strong> length <strong>and</strong> facial<br />

length were not significantly different between<br />

these taxa.<br />

Lieberman (1998), however, incorrectly<br />

measured ASL in the few archaic humans in<br />

which the <strong>cranial</strong> <strong>base</strong> is well preserved. As<br />

shown by Spoor et al. (1999), ASL is not<br />

significantly longer in archaic Homo than in<br />

modern humans, but the angle of the <strong>cranial</strong><br />

<strong>base</strong> (CBA1) is about 15° more extended<br />

in archaic Homo fossils such as Gibraltar,<br />

Monte Circeo, <strong>and</strong> Kabwe than in samples<br />

of Pleistocene <strong>and</strong> recent modern humans<br />

(P 0.05). Consequently, Spoor et al. (1999)<br />

<strong>and</strong> Lieberman (2000) concluded that differences<br />

in <strong>cranial</strong> <strong>base</strong> angle are more likely<br />

to account for facial retraction in modern<br />

humans, as well as for other differences<br />

noted by Lieberman (1998), such as the relatively<br />

shorter pharynx behind the palate.<br />

This hypothesis needs to be tested carefully,<br />

but is explored here in a preliminary fashion<br />

with a geometric morphometric analysis<br />

comparing the shape of the Kabwe cranium<br />

with a large, robust recent H. sapiens (a<br />

male Australian). Figure 17 shows a thinplate<br />

spline transformation of the Australian<br />

into Kabwe (computed using Mor-


158 YEARBOOK OF PHYSICAL ANTHROPOLOGY [Vol. 43, 2000<br />

phometrika, version 0.007, Jeff Walker,<br />

Chicago, IL), <strong>base</strong>d on an initial resistant-fit<br />

superimposition Procrustes analysis of six<br />

<strong>cranial</strong> <strong>base</strong> l<strong>and</strong>marks, seven facial l<strong>and</strong>marks,<br />

<strong>and</strong> five neuro<strong>cranial</strong> l<strong>and</strong>marks. 10<br />

<strong>The</strong> thin-plate spline, which is <strong>base</strong>d on a<br />

Procrustes analysis that geometrically corrects<br />

for most effects of size differences,<br />

shows that Kabwe has a considerably more<br />

projecting face (i.e., between nasion <strong>and</strong> the<br />

foramen caecum) in conjunction with a relatively<br />

more extended <strong>cranial</strong> <strong>base</strong> (14 o ), a<br />

dorsally rotated PM plane relative to the<br />

posterior <strong>cranial</strong> <strong>base</strong>, <strong>and</strong> a relatively<br />

longer pharyngeal space between the maxilla<br />

<strong>and</strong> the foramen magnum. In addition,<br />

when one examines specific, size-corrected<br />

dimensions (st<strong>and</strong>ardized by the geometric<br />

mean of all the interl<strong>and</strong>mark distances),<br />

midfacial length (PM-Na) is 19% longer in<br />

Kabwe than in the Australian modern individual,<br />

but the length of the anterior <strong>cranial</strong><br />

<strong>base</strong> (S-FC) is only 4% shorter. In other<br />

words, Kabwe exhibits a considerably more<br />

projecting face than the H. sapiens specimen<br />

because it has a more extended <strong>cranial</strong><br />

<strong>base</strong> in conjunction with a relatively longer<br />

face. Changes in middle <strong>cranial</strong> fossa shape,<br />

therefore, appear to have had major effects<br />

on facial retraction in human evolution, but<br />

need to be further examined using additional<br />

specimens.<br />

Additional, intriguing evidence that the<br />

<strong>cranial</strong> <strong>base</strong> can play an important role in<br />

facial retraction may be come from studies<br />

of craniofacial growth abnormalities <strong>and</strong><br />

from laboratory studies of mice. Mice that<br />

are homozygous-recessive for the retrognathic<br />

Brachyrrhine (Br) allele are characterized<br />

by a primary growth defect in the<br />

anterior <strong>cranial</strong> <strong>base</strong> (Beechey et al., 1998;<br />

Ma <strong>and</strong> Lozanoff, 1999) that leads to a severely<br />

retrognathic midfacial profile but a<br />

morphologically normal nasal septum <strong>and</strong><br />

face (Lozanoff, 1993; Lozanoff et al., 1994;<br />

10 <strong>The</strong>se l<strong>and</strong>marks are: basion, sella, pituitary point, sphenoidale,<br />

PMp (the anteriormost point on the greater wings of the<br />

sphenoid), foramen caecum, nasion, ANS, PNS, prosthion, maxillary<br />

tuberosities, fronton, orbitale, opisthocranion, the superior-most<br />

point on the vault, fronton, the inferior-most point on<br />

the supratoral sulcus, glabella. Resistant fit superimposition<br />

was used in order to minimize effects of the very different shape<br />

of the neurocranium in archaic Homo vs. modern H. sapiens.<br />

Ma <strong>and</strong> Lozanoff, 1999). <strong>The</strong> extent to<br />

which the morphological differences between<br />

the retrognathic Br mouse <strong>and</strong> controls<br />

is at all similar to the differences evident<br />

between H. sapiens <strong>and</strong> archaic Homo<br />

has yet to be determined. However, these<br />

studies, in conjunction with facial growth<br />

defects caused by chondrogenic growth disorders<br />

such as Crouzon syndrome, Pierre-<br />

Robin syndrome, <strong>and</strong> Down syndrome, highlight<br />

the important role the <strong>cranial</strong> <strong>base</strong><br />

plays in facial growth <strong>and</strong> integration (David<br />

et al., 1989).<br />

Cranial <strong>base</strong> characters in phylogenetic<br />

analyses<br />

One final consideration of interest is<br />

whether it might be profitable to focus on<br />

characters from the <strong>cranial</strong> <strong>base</strong> in taxonomic<br />

<strong>and</strong> phylogenetic analyses of <strong>primate</strong>s.<br />

This possibility has been raised by a<br />

number of authors (e.g., Olson, 1985; Shea,<br />

1985a, 1988; Lieberman et al., 1996; Lieberman,<br />

1997; Strait et al., 1997; Strait, 1998)<br />

for three reasons. First, the <strong>cranial</strong> <strong>base</strong><br />

derives from endochrondral precursors<br />

rather than though intramembranous ossification<br />

processes (endochondral bones are<br />

thought to have more direct genetic influence<br />

in terms of initial shape); second, the<br />

<strong>cranial</strong> <strong>base</strong> is the first part of the skull to<br />

reach adult size <strong>and</strong> shape (Moore <strong>and</strong><br />

Lavelle, 1974); <strong>and</strong> third, the <strong>cranial</strong> <strong>base</strong><br />

may play a greater role in influencing facial<br />

shape than vice versa (see above). Consequently,<br />

one might expect <strong>cranial</strong> <strong>base</strong> characters<br />

to preserve more phylogenetic signal<br />

than facial characters by virtue of being<br />

more heritable <strong>and</strong> less influenced by the<br />

epigenetic responses to external influences<br />

during postnatal <strong>ontogeny</strong> rampant in the<br />

facial skeleton (Herring, 1993).<br />

Expectations aside, there are currently no<br />

data suggesting that the basicranium is actually<br />

a better source of characters than<br />

other regions of the skull for phylogenetic<br />

analyses. In terms of narrow-sense heritability<br />

(h 2 ), basi<strong>cranial</strong>, neuro<strong>cranial</strong>, <strong>and</strong> facial<br />

characters have similar levels of heritability<br />

among <strong>primate</strong>s (Sjøvold, 1984;<br />

Cheverud <strong>and</strong> Buikstra, 1982; Cheverud,<br />

1995). In addition, basi<strong>cranial</strong> variables appear<br />

to be equally well (or poorly) correlated


D.E. Lieberman et al.]<br />

PRIMATE CRANIAL BASE 159<br />

with other dimensions or features in hominoids<br />

(Strait, 1998), <strong>and</strong> in the human skull<br />

(Lieberman et al., 2000). Finally, those few<br />

studies that have focused on <strong>cranial</strong> <strong>base</strong><br />

characters do not yield results substantially<br />

different from analyses that incorporate<br />

other craniodental characters. Lieberman et<br />

al. (1996) found that basi<strong>cranial</strong> <strong>and</strong> vault<br />

characters tend to yield similar cladograms<br />

that differ only slightly from cladograms<br />

<strong>base</strong>d on facial characters. Strait et al.<br />

(1997) did not specifically examine characters<br />

from the <strong>cranial</strong> <strong>base</strong>, but found no<br />

major difference between trees <strong>base</strong>d on<br />

masticatory characters vs. those that were<br />

predominantly neuro<strong>cranial</strong> <strong>and</strong> basi<strong>cranial</strong>.<br />

In their study of <strong>primate</strong> higher taxonomic<br />

relationships, Ross et al. (1998) found<br />

the <strong>cranial</strong> data, consisting primarily of basi<strong>cranial</strong><br />

traits, to yield trees similar to<br />

those produced by the dental data, although<br />

with better resolution at older nodes. Thus,<br />

whether the basicranium is a better source<br />

of phylogenetic data remains open to question.<br />

<strong>The</strong> <strong>cranial</strong> <strong>base</strong> may be a good place to<br />

look for reliable characters if researchers<br />

are to focus on characters that describe developmental<br />

processes or events (e.g., flexion<br />

vs. extension at the spheno-occipital<br />

synchondrosis) rather than using characters<br />

that solely describe morphological variation<br />

(e.g., the angle of the whole <strong>cranial</strong><br />

<strong>base</strong>) (Hall, 1994; Lieberman, 1999). This<br />

hypothesis, however, has yet to be tested.<br />

<strong>The</strong> above studies, however, raise another<br />

key issue relevant to phylogenetic studies:<br />

the problem of independence. Cladistic<br />

analyses explicitly require one to use independent<br />

characters to avoid problems of<br />

convergent <strong>and</strong> correlated characters incorrectly<br />

biasing the outcome of any parsimony<br />

analysis. Yet the above studies demonstrate<br />

the existence of multiple <strong>and</strong> complicated<br />

interactions between the <strong>cranial</strong> <strong>base</strong> <strong>and</strong><br />

the neurocranium <strong>and</strong> between the <strong>cranial</strong><br />

<strong>base</strong> <strong>and</strong> the face. For example, variations<br />

in orbit size <strong>and</strong> orientation are linked to<br />

variations in the angle of the <strong>cranial</strong> <strong>base</strong>,<br />

brain size, <strong>and</strong> the orientation of the face<br />

relative to the rest of the skull. Many of<br />

these features have been treated as independent<br />

characters in recent cladistic analyses<br />

(e.g., Strait et al, 1997), but their effect<br />

on the results has yet to be determined.<br />

Further study of these interactions is<br />

needed to improve the reliability of phylogenetic<br />

analyses, <strong>and</strong> again highlight the need<br />

to consider morphological characters in<br />

terms of their generative processes (e.g.,<br />

Gould, 1977; Cheverud, 1982; Shea, 1985b;<br />

Hall, 1994; Lieberman, 1999).<br />

CONCLUSIONS<br />

Since the last major reviews of <strong>cranial</strong><br />

<strong>base</strong> anatomy in <strong>primate</strong>s (Scott, 1958;<br />

Moore <strong>and</strong> Lavelle, 1974; Sirianni <strong>and</strong><br />

Swindler, 1979), there has been a tremendous<br />

increase in our knowledge of chondro<strong>cranial</strong><br />

embryology, the patterns <strong>and</strong> processes<br />

of basi<strong>cranial</strong> growth, <strong>and</strong> the nature<br />

of basi<strong>cranial</strong> variation across <strong>primate</strong>s.<br />

Major advances include details of the morphogenetic<br />

independence between the prechordal<br />

<strong>and</strong> postchordal portions of the<br />

chondrocranium; comparative data on the<br />

relative importance of brain size, orbital orientation,<br />

facial orientation, <strong>and</strong> posture as<br />

factors that account for variation in <strong>cranial</strong><br />

<strong>base</strong> angulation; <strong>and</strong> ontogenetic <strong>and</strong> comparative<br />

data on the structural relationships<br />

between the anterior <strong>cranial</strong> <strong>base</strong> <strong>and</strong><br />

upper face in haplorhines vs. strepsirhines,<br />

<strong>and</strong> their influences on facial form. Despite<br />

these advances, many aspects of <strong>cranial</strong><br />

<strong>base</strong> growth, variation, <strong>and</strong> <strong>function</strong> remain<br />

poorly understood. For example, we do not<br />

know what ontogenetic processes govern<br />

flexion <strong>and</strong> extension of the <strong>cranial</strong> <strong>base</strong>, or<br />

which synchondroses are active in <strong>cranial</strong><br />

<strong>base</strong> elongation vs. angulation in humans<br />

<strong>and</strong> other <strong>primate</strong>s. Future research needs<br />

to be aimed at studying these processes as<br />

they relate to <strong>cranial</strong> shape, <strong>function</strong>, <strong>and</strong><br />

evolution.<br />

To conclude, we highlight two important<br />

practical <strong>and</strong> theoretical issues which we<br />

believe merit special consideration, <strong>and</strong><br />

which promise to further our underst<strong>and</strong>ing<br />

of craniofacial growth <strong>and</strong> variation. First,<br />

what are the major factors that generate<br />

variation in the <strong>cranial</strong> <strong>base</strong> among <strong>primate</strong>s?<br />

Second, to what extent does the <strong>cranial</strong><br />

<strong>base</strong> <strong>function</strong> to coordinate these factors<br />

within the craniofacial complex during<br />

growth <strong>and</strong> development? As noted above,<br />

these questions need to be addressed using


160 YEARBOOK OF PHYSICAL ANTHROPOLOGY [Vol. 43, 2000<br />

two approaches. Ontogenetic studies are<br />

crucial for testing hypotheses about the generation<br />

of morphological covariation patterns,<br />

<strong>and</strong> comparative studies are important<br />

for probing the extent to which<br />

interspecific patterns of morphological evolution<br />

are epiphenomena of developmental<br />

processes.<br />

What major factors generate variation<br />

in the <strong>cranial</strong> <strong>base</strong>?<br />

<strong>The</strong> studies summarized above suggest<br />

that relative brain size, particularly relative<br />

to basi<strong>cranial</strong> length, is an important determinant<br />

of the degree of basi<strong>cranial</strong> angulation.<br />

However, the effects of (especially<br />

prenatal) brain shape on basi<strong>cranial</strong><br />

morphology have yet to be thoroughly investigated.<br />

Further experimental <strong>and</strong> comparative<br />

morphological research is needed to<br />

relate soft-tissue <strong>and</strong> bony morphology in<br />

the <strong>primate</strong> head. Despite a long tradition<br />

suggesting links between basi<strong>cranial</strong> morphology<br />

<strong>and</strong> head, neck, or body posture,<br />

there is currently little empirical support<br />

for the hypothesis that these factors are directly<br />

related to variation in the <strong>cranial</strong><br />

<strong>base</strong>, especially angulation in the midsagittal<br />

plane. Better data are also needed on<br />

head <strong>and</strong> neck posture during locomotion<br />

before locomotor-related <strong>cranial</strong> adaptations<br />

can be definitively identified. However,<br />

connections between facial orientation<br />

<strong>and</strong> basi<strong>cranial</strong> morphology on the one h<strong>and</strong><br />

<strong>and</strong> head posture on the other leave open<br />

the possibility of an indirect link between<br />

basi<strong>cranial</strong> angle <strong>and</strong> head posture.<br />

What role does the <strong>cranial</strong> <strong>base</strong> play in<br />

craniofacial integration?<br />

Because many variables influence <strong>cranial</strong><br />

<strong>base</strong> shape, it follows that these variables<br />

also influence other aspects of <strong>cranial</strong> shape<br />

via the <strong>cranial</strong> <strong>base</strong>. Consequently, a key<br />

issue that emerges repeatedly in discussions<br />

of the role of the basicranium in <strong>cranial</strong><br />

development, growth, <strong>and</strong> evolution is<br />

integration. Integration, which is defined<br />

here as “the association of elements through<br />

a set of causal mechanisms so that change<br />

in one element is reflected by change in<br />

another” (Smith, 1996, p. 70), thereby generating<br />

a pattern of significant covariation<br />

(see also Olson <strong>and</strong> Miller, 1958; Cheverud,<br />

1982; Zelditch, 1988), 11 is an important issue<br />

because of the many genetic, developmental,<br />

<strong>and</strong> <strong>function</strong>al interactions that occur<br />

between the basicranium <strong>and</strong> its<br />

neighboring anatomical components. As described<br />

above, the <strong>cranial</strong> <strong>base</strong> is likely to<br />

directly interact, developmentally <strong>and</strong> <strong>function</strong>ally,<br />

with various adjoining skeletal,<br />

muscular, <strong>and</strong> neurosensory complexes,<br />

most notably the brain, the orbits, the ethmomaxillary<br />

complex, <strong>and</strong> the neck. <strong>The</strong>se<br />

“units,” in turn, have direct <strong>and</strong> indirect<br />

interactions with other putative units such<br />

as the oropharynx, nasopharynx, m<strong>and</strong>ible,<br />

<strong>and</strong> maxillary arches.<br />

But does the <strong>cranial</strong> <strong>base</strong> play an active or<br />

a passive role in integrating <strong>cranial</strong> shape<br />

among these disparate units, <strong>and</strong> how much<br />

integration actually occurs? On an intuitive<br />

level, there are several reasons to suggest<br />

that the <strong>cranial</strong> <strong>base</strong> acts in part as a structural<br />

“interface” during growth between the<br />

brain <strong>and</strong> the face, <strong>and</strong> between the head<br />

<strong>and</strong> the neck. In many regions, the basicranium<br />

serves as the actual structural boundary<br />

between disparate components of the<br />

skull: the floor of the anterior <strong>cranial</strong> fossa<br />

is the roof of the orbits; the back of the<br />

midface is the front of the middle <strong>cranial</strong><br />

fossa; <strong>and</strong> the posterior <strong>cranial</strong> <strong>base</strong> is the<br />

posterior roof of the oropharynx. Yet, in<br />

spite of these obvious relationships, it is<br />

difficult to define or assess quantitative hypotheses<br />

about craniofacial integration because<br />

we know so little about the extent <strong>and</strong><br />

nature of the numerous interactions that<br />

presumably occur between <strong>and</strong> among regions<br />

of the <strong>cranial</strong> <strong>base</strong> <strong>and</strong> other parts of<br />

the skull. What are the actual units in the<br />

<strong>cranial</strong> <strong>base</strong> <strong>and</strong> skull that interact, <strong>and</strong><br />

what regulates their interrelations <strong>and</strong>, especially,<br />

interactions? In other words, the<br />

appropriate null hypothesis to be tested is<br />

that the <strong>cranial</strong> <strong>base</strong>, while associated with<br />

variation in other parts of the skull, plays<br />

11 Note that some researchers (most notably Olson <strong>and</strong> Miller,<br />

1958) define morphological integration solely as a multivariate<br />

pattern of covariation with regard to some a priori biological<br />

hypothesis (thus recognizing the importance of generative processes),<br />

whereas others (e.g., Cheverud, 1996) define integration<br />

more generally as either a pattern or a process that refers to<br />

“connections or relationships among morphological elements.”


D.E. Lieberman et al.]<br />

PRIMATE CRANIAL BASE 161<br />

no more of an integrative role than any<br />

other part of the skull (e.g., the m<strong>and</strong>ible,<br />

the brain).<br />

In order to explore the basicranium’s role<br />

in craniofacial integration, it is first necessary<br />

to decide how to test hypotheses about<br />

the many different processes through which<br />

integration occurs, <strong>and</strong> the many structural<br />

<strong>and</strong> <strong>function</strong>al levels at which integration is<br />

evident. Readers interested in this huge <strong>and</strong><br />

complex topic (which is too large to review<br />

thoroughly here) should consult the recent<br />

review by Chernoff <strong>and</strong> Magwene (1999). As<br />

a first-order analysis, it is useful to test<br />

hypotheses about developmental integration<br />

by considering the processes by which<br />

different components of a system influence<br />

one another. Thus, in the context of <strong>cranial</strong><br />

<strong>base</strong> growth within the skull, integration<br />

can occur through the direct inductive<br />

<strong>and</strong>/or mechanical effects of neighboring tissue-tissue<br />

interactions (see Moss, 1997);<br />

through secondary effects of growth in one<br />

region causing changes in the positional relationships<br />

among bones in other regions;<br />

<strong>and</strong> genetically through single genes which<br />

have effects on multiple regions, or through<br />

the coordinated action of multiple genes via<br />

pleiotropy <strong>and</strong> linkage (Atchley <strong>and</strong> Hall,<br />

1991; Cheverud, 1982, 1995, 1996; Zelditch<br />

<strong>and</strong> Fink, 1995, Zelditch <strong>and</strong> Fink, 1996;<br />

Chernoff <strong>and</strong> Magwene, 1999). Of course,<br />

the strength of such ontogenetic interactions<br />

can vary due to changes in the timing<br />

of developmental events, something of great<br />

importance for the analysis of heterochrony,<br />

as well as due to allometric (size-correlated,<br />

size-required) factors. Note that we consider<br />

here integration solely in terms of developmental<br />

events <strong>and</strong> processes (e.g., epigenetic<br />

responses to mechanical loading, displacement<br />

due to cell division), but it is<br />

useful to recognize that developmental integration<br />

leads to, <strong>and</strong> is sometimes caused<br />

by, structural <strong>and</strong> <strong>function</strong>al integration.<br />

Another problem with assessing any integrative<br />

role of the <strong>cranial</strong> <strong>base</strong> within the<br />

skull is the lack of widely accepted criteria<br />

for testing hypotheses of integration. Following<br />

Olson <strong>and</strong> Miller (1958) <strong>and</strong><br />

Chernoff <strong>and</strong> Magwene (1999), we apply<br />

three nonexclusive criteria (correlation, constraint,<br />

<strong>and</strong> ontogenetic sequence) in a preliminary<br />

fashion to solely phenotypic variation<br />

in the <strong>cranial</strong> <strong>base</strong>, using the data<br />

presented above.<br />

Correlation. Integration is most basically<br />

revealed by complex patterns of correlation<br />

<strong>and</strong> covariation which indicate a lack<br />

of independence among variables (see Cheverud,<br />

1995, 1996), <strong>and</strong> which can be recognized<br />

a posteriori by comparing theoretically<br />

<strong>and</strong> empirically derived correlation<br />

matrices (Cheverud, 1982; Shea, 1985b;<br />

Zelditch, 1987, 1988; Cheverud et al., 1989;<br />

Wagner, 1989). Such studies have yet to be<br />

carried out for models that explicitly focus<br />

on the <strong>primate</strong> <strong>cranial</strong> <strong>base</strong> (but see Cheverud,<br />

1995). However, as noted above, a<br />

high degree of covariation is frequently evident<br />

in interspecific analyses of variation<br />

in the <strong>cranial</strong> <strong>base</strong> <strong>and</strong> other parts of the<br />

skull. Perhaps the most obvious example of<br />

this phenomenon is the percentage of variation<br />

in <strong>cranial</strong> <strong>base</strong> angle accounted for by<br />

factors such as brain volume, basi<strong>cranial</strong><br />

length, facial angle, <strong>and</strong> posture. IRE1,<br />

which appears to be the dominant factor<br />

that influences <strong>cranial</strong> <strong>base</strong> angle in <strong>primate</strong>s,<br />

explains 58% of the variation in<br />

CBA1 <strong>and</strong> 38% of the variation in CBA4 (see<br />

Table 5). Furthermore, according to Ross<br />

<strong>and</strong> Ravosa (1993, their Table 2), orbital<br />

axis orientation explains 41% of the variation<br />

in CBA4 among <strong>primate</strong>s, while facial<br />

orientation explains 22% of the variation in<br />

CBA4 among <strong>primate</strong>s. <strong>The</strong> partial correlation<br />

analysis of Strait <strong>and</strong> Ross (1999) (discussed<br />

above) found that IRE explained 36%<br />

of the variation in CBA1 when the effects of<br />

orbital axis <strong>and</strong> head-neck angle were factored<br />

out. Further evidence for this sort of<br />

complex pattern of covariation is documented<br />

among humans between CBA, <strong>cranial</strong><br />

<strong>base</strong> length, <strong>cranial</strong> <strong>base</strong> width, <strong>and</strong><br />

brain volume (Lieberman et al., 2000).<br />

<strong>The</strong>se results are indicative (but not proof)<br />

of a pattern whereby multiple factors combine<br />

to influence CBA in such a way that<br />

variation in CBA itself may play some role<br />

in modulating the interactions among different,<br />

spatially separated components of<br />

the cranium. This hypothesis, however, requires<br />

further testing with comparative ontogenetic<br />

data.


162 YEARBOOK OF PHYSICAL ANTHROPOLOGY [Vol. 43, 2000<br />

Constraint. Another expected outcome of<br />

integration is constraint, which is defined<br />

most generally as a restriction or limitation<br />

on variation. Regardless of their causes,<br />

which can be phylogenetic, <strong>function</strong>al, developmental,<br />

or structural (Alberch, 1985;<br />

Maynard Smith et al., 1985), constraints are<br />

most basically evident in patterns of invariance.<br />

<strong>The</strong>re has not been much work on<br />

phenotypic constraint in the <strong>cranial</strong> <strong>base</strong>,<br />

but several examples suggest it deserves<br />

further research. One source of evidence for<br />

constraint is revealed by allometry, which<br />

measures size-related conservation of shape<br />

(keeping in mind that size-correlated patterns<br />

due to epigenetic <strong>and</strong> perhaps genetic<br />

factors differ from size-required patterns<br />

that have a primarily <strong>function</strong>al basis). As<br />

Table 4 (see also Strait, 1999; McCarthy,<br />

2001) illustrates, there are numerous,<br />

strongly correlated ontogenetic scaling relationships<br />

across <strong>primate</strong>s between various<br />

components of the <strong>cranial</strong> <strong>base</strong>, the volume<br />

of neural regions, <strong>and</strong> facial dimensions. Integration<br />

between the brain <strong>and</strong> face via the<br />

<strong>cranial</strong> <strong>base</strong> is implied by the fact that many<br />

scaling relationships between contiguous<br />

anatomical units (e.g., the noncortical brain<br />

<strong>and</strong> the posterior <strong>cranial</strong> <strong>base</strong>) result in additional<br />

strongly correlated scaling relationships<br />

between noncontiguous anatomical<br />

units, such as those between brain volume<br />

<strong>and</strong> facial size. <strong>The</strong>se scaling relationships<br />

require more study.<br />

Another potential source of evidence regarding<br />

the presence of a constraint are patterns<br />

of angular invariance. One important<br />

example is the apparently invariant 90° angle<br />

between the PM plane <strong>and</strong> the NHA,<br />

<strong>and</strong> the limited variability that results from<br />

this relationship on the angle between the<br />

PM plane <strong>and</strong> both the planum sphenoideum<br />

<strong>and</strong> the anterior <strong>cranial</strong> <strong>base</strong> (S-FC)<br />

in anthropoids (McCarthy <strong>and</strong> Lieberman,<br />

2001). Other less secure examples of invariance<br />

may include the relationship between<br />

cribriform plate orientation <strong>and</strong> facial orientation<br />

(Ravosa <strong>and</strong> Shea, 1994), <strong>and</strong> the<br />

near 45° angle between the external auditory<br />

meatus, the maxillary tuberosities, <strong>and</strong><br />

the midpoint of the orbital aperture (Bromage,<br />

1992). Further research, however, is<br />

needed on the extent to which invariant angles<br />

<strong>and</strong> spatial relationships occur in the<br />

skull, <strong>and</strong> additional research is needed to<br />

assess the developmental, structural, <strong>and</strong><br />

<strong>function</strong>al <strong>base</strong>s of these relationships, <strong>and</strong><br />

whether they reflect constraints that result<br />

from integration.<br />

Ontogenetic sequence. Finally, hypotheses<br />

of phenotypic integration may sometimes<br />

be inferred or tested by examining<br />

ontogenetic sequences during normal<br />

growth <strong>and</strong> in the context of controlled experiments.<br />

Ontogenetic data allow one to<br />

test hypotheses of integration by examining<br />

the structural relationships between one<br />

variable <strong>and</strong> another during growth (e.g.,<br />

heterochrony, heterotopy), <strong>and</strong> to compare<br />

the pattern <strong>and</strong> timing of developmental<br />

events. So far, there have been few attempts<br />

to examine hypotheses of integration in the<br />

<strong>cranial</strong> <strong>base</strong> using ontogenetic data (especially<br />

from embryonic stages), but a few examples<br />

indicate that the sequence of interactions<br />

between the <strong>cranial</strong> <strong>base</strong>, the face,<br />

<strong>and</strong> the brain are complex <strong>and</strong> multiphasic,<br />

with the <strong>cranial</strong> <strong>base</strong> mediating various interactions<br />

between the face <strong>and</strong> the brain.<br />

For example, the prenatal human <strong>cranial</strong><br />

<strong>base</strong> initially flexes, then remains stable,<br />

<strong>and</strong> then extends, all during periods of rapid<br />

neural growth (Jeffery, 1999). Postnatally,<br />

the nonhuman <strong>primate</strong> <strong>cranial</strong> <strong>base</strong> (best<br />

studied in Pan <strong>and</strong> Macaca) appears to extend<br />

slightly during the period of neural<br />

growth, <strong>and</strong> subsequently extends more<br />

rapidly <strong>and</strong> for a long time as the face continues<br />

to grow. In contrast, the human <strong>cranial</strong><br />

<strong>base</strong> flexes rapidly during the first few<br />

years of brain growth, <strong>and</strong> subsequently remains<br />

stable. <strong>The</strong>se contrasting sequences<br />

imply multiple interactions between the<br />

brain <strong>and</strong> the face via the <strong>cranial</strong> <strong>base</strong>, but<br />

have yet to be resolved in terms of the actual<br />

processes that cause flexion <strong>and</strong> extension<br />

at specific locations during different periods<br />

of growth. One hypothesis, which remains<br />

to be tested, is that the <strong>cranial</strong> <strong>base</strong> <strong>function</strong>s<br />

to accommodate <strong>and</strong> perhaps coordinate<br />

these different aspects of growth. Controlled<br />

experimental studies, which can<br />

potentially isolate local <strong>and</strong> regional effects<br />

of specific growth stimuli on the <strong>cranial</strong> <strong>base</strong><br />

<strong>and</strong> the rest of the skull, are a promising


D.E. Lieberman et al.]<br />

PRIMATE CRANIAL BASE 163<br />

avenue for future research on this problem<br />

(see Sarnat, 1982; Bütow, 1990; Reidenberg<br />

<strong>and</strong> Laitman, 1991).<br />

Future research<br />

Although the <strong>cranial</strong> <strong>base</strong> does appear to<br />

play some integrative role in the cranium,<br />

we have only a vague, incomplete picture of<br />

how this integration occurs, <strong>and</strong> how much<br />

of a role it plays in influencing various aspects<br />

of craniofacial form <strong>and</strong> <strong>function</strong>, <strong>and</strong><br />

how these processes relate to evolutionary<br />

shifts in the <strong>primate</strong> <strong>cranial</strong> <strong>base</strong>. More research<br />

is needed to isolate <strong>and</strong> define the<br />

actual morphogenetic units which interact,<br />

to identify <strong>and</strong> quantify their direct <strong>and</strong> indirect<br />

interactions, <strong>and</strong> to underst<strong>and</strong> the<br />

processes by which they interact. <strong>The</strong>se<br />

goals may be accomplished by combining at<br />

least four approaches. First, we need to<br />

gather more three-dimensional data on <strong>cranial</strong><br />

<strong>base</strong> variation in ontogenetic samples<br />

at all stages of growth <strong>and</strong> development (fetal<br />

to adult) among different species. <strong>The</strong><br />

<strong>cranial</strong> <strong>base</strong> <strong>and</strong> the rest of the skull comprise<br />

a complex three-dimensional structure<br />

whose internal <strong>and</strong> external structures differ<br />

substantially, yet most studies of <strong>cranial</strong><br />

<strong>base</strong> variation so far have used external<br />

l<strong>and</strong>marks <strong>and</strong>/or two-dimensional radiographic<br />

analysis of midsagittal l<strong>and</strong>marks<br />

(or nonmidsagittal l<strong>and</strong>marks projected into<br />

the midsagittal plane). Second, more data<br />

are needed on the developmental mechanisms<br />

which generate variation in the <strong>cranial</strong><br />

<strong>base</strong>, <strong>and</strong> which regulate interactions<br />

among components of the skull. Experimental,<br />

histological, <strong>and</strong> other kinds of developmental<br />

information will be useful in this<br />

regard, because observed morphological<br />

patterns are potentially generated by different<br />

genetic <strong>and</strong> epigenetic processes. Third,<br />

more data are needed on the genetic <strong>base</strong>s<br />

for variation in <strong>cranial</strong> <strong>base</strong> growth <strong>and</strong><br />

form. Lastly, more detailed interspecific<br />

analyses are needed to extend the evolutionary<br />

implications of ontogenetic tests of hypotheses<br />

about interactions <strong>and</strong> relationships<br />

among morphogenetic units of the<br />

skull. Such future research on the <strong>cranial</strong><br />

<strong>base</strong> should provide interesting <strong>and</strong> valuable<br />

insights on other aspects of craniofacial<br />

growth, <strong>function</strong>, <strong>and</strong> evolution in <strong>primate</strong>s<br />

<strong>and</strong> other mammals.<br />

ACKNOWLEDGMENTS<br />

We thank N. Jeffery, W. Jungers, R. Mc-<br />

Carthy, C. Ruff, B. Shea, F. Spoor, <strong>and</strong> two<br />

anonymous reviewers for their comments<br />

<strong>and</strong> assistance.<br />

LITERATURE CITED<br />

Adams LM, Moore WJ. 1975. A biomechanical appraisal<br />

of some skeletal features associated with head balance<br />

<strong>and</strong> posture in the Hominoidea. Acta Anat<br />

(Basel) 92:580–594.<br />

Aiello L, Dean C. 1990. An introduction to human evolutionary<br />

anatomy. San Diego: Academic Press.<br />

Alberch P. 1985. Developmental constraints: why St.<br />

Bernards often have an extra digit <strong>and</strong> poodles never<br />

do. Am Nat 126:430–433.<br />

Allman J. 1977. Evolution of the visual system in early<br />

<strong>primate</strong>s. In: Sprague JM, Epstein JM, editors.<br />

Progress in psychobiology, physiology, <strong>and</strong> psychology.<br />

Volume 7. New York: Academic Press. p 1–53.<br />

Allman J. 1982. Reconstructing the evolution of the<br />

brain in <strong>primate</strong>s through the use of comparative<br />

neurophysiological <strong>and</strong> neuroanatomical data. In:<br />

Armstrong E, Falk D, editors. Primate brain evolution.<br />

New York: Plenum Press. p 13–28.<br />

Allman JM, Kaas JH. 1974. <strong>The</strong> organization of the<br />

second visual area (V II) in the owl monkey: a second<br />

order transformation of the visual hemifield. Brain<br />

Res 76:247–265.<br />

Anagastopolou S, Karamaliki DD, Spyropoulos MN.<br />

1988. Observations on the growth <strong>and</strong> orientation of<br />

the anterior <strong>cranial</strong> <strong>base</strong> in the human foetus. Eur<br />

J Orthod 10:143–148.<br />

Angst R. 1967. Beitrag zum formw<strong>and</strong>el des craniums<br />

der ponginen. Z Morphol Anthropol 58:109–151.<br />

Antón SC. 1989. Intentional <strong>cranial</strong> vault deformation<br />

<strong>and</strong> induced changes in the <strong>cranial</strong> <strong>base</strong> <strong>and</strong> face.<br />

Am J Phys Anthropol 79:253–267.<br />

Antón SC. 1994. Mechanical <strong>and</strong> other perspectives on<br />

Ne<strong>and</strong>ertal craniofacial morphology. In: Corruccini<br />

RS, Ciochon RL, editors. Integrative paths to the<br />

past: palaeoanthropological advances in honor of F.<br />

Clark Howell. Englewood Cliffs, NJ: Prentice Hall. p<br />

677–695.<br />

Ashton EH. 1952. Age changes in the basi<strong>cranial</strong> axis of<br />

the anthropoidea. Proc Zool Soc Lond [Biol] 129:61–<br />

74.<br />

Ashton EH, Spence TF. 1958. Age changes in the <strong>cranial</strong><br />

capacity <strong>and</strong> foramen magnum of hominoids.<br />

Proc Zool Soc Lond [Biol] 130:169–181.<br />

Ashton EH, Zuckerman S. 1952. Age changes in the<br />

position of the occipital condyles in the chimpanzee<br />

<strong>and</strong> gorilla. Am J Phys Anthropol 10:277–288.<br />

Ashton EH, Zuckerman S. 1956. Age changes in the<br />

position of the foramen magnum in hominoids. Proc<br />

Zool Soc Lond [Biol] 126:315–325.<br />

Atchley WR, Hall BK. 1991. A model for development<br />

<strong>and</strong> evolution of complex morphological structures.<br />

Biol Rev 66:101–157.<br />

Babler WJ. 1989. Relationship of altered <strong>cranial</strong> suture<br />

growth to the <strong>cranial</strong> <strong>base</strong> <strong>and</strong> midface. In: Persing<br />

JA, Edgerton MT, Jane JA, editors. Scientific foundations<br />

<strong>and</strong> surgical treatment of craniosynostosis. Baltimore:<br />

Williams <strong>and</strong> Wilkins. p 87–95.<br />

Baer T, Gore JC, Gracco LC, Nye PW. 1991. Analysis of<br />

vocal tract shape <strong>and</strong> dimensions using magnetic res-


164 YEARBOOK OF PHYSICAL ANTHROPOLOGY [Vol. 43, 2000<br />

onance imaging: vowels. J Acoust Soc Am 90:799–<br />

828.<br />

Barlow HB. 1986. Why have multiple cortical areas?<br />

Vision Res 26:81–90.<br />

Barton RA. 1998. Visual specialization <strong>and</strong> brain evolution<br />

in <strong>primate</strong>s. Proc R Soc Lond [Biol] 265:1933–<br />

1937.<br />

Baume JL. 1957. A biologist looks at the sella point. Eur<br />

Orthod Soc Trans 33:150–159.<br />

Beard KC. 1993. Phylogenetic systematics of the Primatomorpha,<br />

with special reference to Dermoptera.<br />

In: Szalay FS, Novacek MJ, McKenna MC, editors.<br />

Mammal phylogeny: placentals. New York: Springer-<br />

Verlag. p 129–150.<br />

Beckman ME, Jung T-P, Lee S-H, de Jong K, Krishnamurthy<br />

AK, Ahalt SC, Cohen KB, Collins MJ.<br />

1995. Variability in the production of quantal vowels<br />

revisited. J Acoust Soc Am 97:471–489.<br />

Beechey C, Boyd Y, Searle AG. 1998. Brachyrrhine, Br,<br />

a mouse craniofacial mutant maps to distal mouse<br />

chromosome 17 <strong>and</strong> is a c<strong>and</strong>idate model for midline<br />

cleft syndrome. Mouse Genome 95:692–694.<br />

Biegert J. 1957. Der Formw<strong>and</strong>el der Primatenschädels<br />

und seine Beziehungen zur ontogenetischen Entwicklung<br />

und den phylogenetischen. Gegenbaurs Morphol<br />

Jahrb 98:77–199.<br />

Biegert J. 1963. <strong>The</strong> evaluation of characteristics of the<br />

skull, h<strong>and</strong>s <strong>and</strong> feet for <strong>primate</strong> taxonomy. In: Washburn<br />

SL, editor. Classification <strong>and</strong> human evolution.<br />

Chicago: Aldine. p 116–145.<br />

Björk A. 1955. Cranial <strong>base</strong> development. Am J Orthod<br />

41:198–225.<br />

Bolk L. 1909. On the position <strong>and</strong> displacement of the<br />

foramen magnum in the Primates. Verh Akad Wet<br />

Amst 12:362–377.<br />

Bolk L. 1910. On the slope of the foramen magnum in<br />

Primates. Verh Akad Wet Amst 12:525–534.<br />

Bolk L. 1926. Das Problem der Menschwerdung. Jena:<br />

Gustav Fischer.<br />

Bookstein F. 1991. Morphometric tools for l<strong>and</strong>mark<br />

data: geometry <strong>and</strong> biology. New York: John Wiley<br />

<strong>and</strong> Sons.<br />

Bosma JF, editor. 1976. Development of the basicranium.<br />

Bethesda, MD: National Institutes of Health.<br />

Braga J, Crubezy E, Elyaqtine M. 1999. <strong>The</strong> posterior<br />

border of the sphenoid greater wing <strong>and</strong> its phylogenetic<br />

usefulness in human evolution. Am J Phys Anthropol<br />

107:387–399.<br />

Brodie AG. 1941. On the growth pattern of the human<br />

head from the third month to the eighth year of life.<br />

Am J Anat 68:208–262.<br />

Brodie AG. 1953. Late growth changes in the human<br />

face. Angle Orthod 23:146–157.<br />

Bromage TG. 1992. <strong>The</strong> <strong>ontogeny</strong> of Pan troglodytes<br />

craniofacial architectural relationships <strong>and</strong> implications<br />

for early hominids. J Hum Evol 23:235–251.<br />

Butler AB, Hodos W. 1996. Comparative vertebrate<br />

neuroanatomy, evolution <strong>and</strong> adaptation. New York:<br />

Wiley-Liss.<br />

Bütow K-W. 1990. Craniofacial growth disturbance after<br />

skull <strong>base</strong> <strong>and</strong> associated suture synostoses in the<br />

newborn Chacma baboon: a preliminary report. Cleft<br />

Palate J 27:241–252.<br />

Cameron J. 1924. <strong>The</strong> cranio-facial axis of Huxley. Part<br />

I. Embryological considerations. Trans R Soc Can 18:<br />

115–123.<br />

Cameron J. 1927. <strong>The</strong> main angle of <strong>cranial</strong> flexion (the<br />

nasion-pituitary-basion angle). Am J Phys Anthropol<br />

10:275–279.<br />

Cameron J. 1930. Craniometric memoirs. No. II. <strong>The</strong><br />

human <strong>and</strong> comparative anatomy of Cameron’s<br />

craniofacial axis. J Anat 64:324–336.<br />

Carre R, Lindblom B, MacNeilage P. 1994. Acoustic<br />

factors in the evolution of the human vocal tract. J<br />

Acoust Soc Am 95:2924.<br />

Cartmill M. 1970. <strong>The</strong> orbits of arboreal mammals.<br />

Ph.D. dissertation, <strong>University</strong> of Chicago.<br />

Cartmill M. 1972. Arboreal adaptations <strong>and</strong> the origin<br />

of the order Primates. In: Tuttle RH, editor. <strong>The</strong> <strong>function</strong>al<br />

<strong>and</strong> evolutionary biology of <strong>primate</strong>s. Chicago:<br />

Aldine. p 97–122.<br />

Cartmill M. 1974. Rethinking <strong>primate</strong> origins. Science<br />

184:436–443.<br />

Cartmill M. 1980. Morphology, <strong>function</strong>, <strong>and</strong> evolution<br />

of the anthropoid postorbital septum. In: Ciochon RL<br />

Chiarelli AB, editors. Evolutionary biology of the New<br />

World monkeys <strong>and</strong> continental drift. New York: Plenum<br />

Press. p 243–274.<br />

Cartmill M. 1992. New views on <strong>primate</strong> origins. Evol<br />

Anthropol 1:105–111.<br />

Chan L-K. 1991. <strong>The</strong> deduction of the level of the human<br />

larynx from bony l<strong>and</strong>marks: its relevance to the<br />

evolution of speech. Hum Evol 6:249–261.<br />

Cherniak C. 1995. Neural component placement.<br />

Trends Neurosci 18:522–527.<br />

Chernoff B, Magwene P. 1999. Morphological integration:<br />

forty years later. In: Olson EC, Miller RL, editors.<br />

Morphological integration. Chicago: <strong>University</strong><br />

of Chicago Press. p 319–348.<br />

Cheverud JM. 1982. Phenotypic, genetic, <strong>and</strong> environmental<br />

morphological integration in the cranium.<br />

Evolution 36:499–516.<br />

Cheverud JM. 1995. Morphological integration in the<br />

saddle back tamarin (Saguinas fuscicollis) cranium.<br />

Am Nat 145:63–89.<br />

Cheverud JM. 1996. Developmental integration <strong>and</strong> the<br />

evolution of pleiotropy. Am Zool 36:44–50.<br />

Cheverud JM, Buikstra J. 1982. Quantitative genetics<br />

of skeletal non-metric traits in the rhesus macaques<br />

on Cayo Santiago III. Relative heritability of skeletal<br />

metric <strong>and</strong> non-metric traits. Am J Phys Anthropol<br />

59:151–155.<br />

Cheverud JM, Richtsmeier J. 1986. Finite element scaling<br />

applied to sexual dimorphism in rhesus macaque<br />

(Macaca mulatta) facial growth. Syst Zool 35:381–<br />

399.<br />

Cheverud JM, Dow MM, Leutenegger W. 1985. <strong>The</strong><br />

quantitative assessment of phylogenetic constraints<br />

in comparative analyses: sexual dimorphism in body<br />

weight among <strong>primate</strong>s. Evolution 39:1335–1351.<br />

Cheverud JM, Wagner GP, Dow MM. 1989. Methods for<br />

the comparative analysis of variation patterns. Syst<br />

Zool 38:201–213.<br />

Cheverud JM, Kohn LAP, Konigsberg LW, Leigh SR.<br />

1992. <strong>The</strong> effects of fronto-occipital artifical <strong>cranial</strong><br />

vault modification on the <strong>cranial</strong> <strong>base</strong> <strong>and</strong> face. Am J<br />

Phys Anthropol 88:323–345.<br />

Couly GF, Coltey PM, Le Douarin NM. 1993. <strong>The</strong> triple<br />

origin of skull in higher vertebrates. Development<br />

117:409–429.<br />

Cousin RP, Fenart R, Deblock R. 1981. Variations ontogénétiques<br />

des angles basicraniens et faciaux. Bull<br />

Mem Soc Anthropol Paris 8:189–212.<br />

Covert HH, Hamrick MW. 1993. Description of new<br />

skeletal remains of the early Eocene anaptomorphine<br />

<strong>primate</strong> Absarokius (Omomyidae) <strong>and</strong> a discussion<br />

about its adaptive profile. J Hum Evol 25:351–362.<br />

Cramer DL. 1977. Craniofacial morphology of Pan paniscus:<br />

a morphometric <strong>and</strong> evolutionary appraisal.<br />

Contrib Primatol 10:1–64.<br />

Crelin E. 1969. Anatomy of the newborn: an atlas. Philadelphia:<br />

Lea <strong>and</strong> Febiger.<br />

Crelin ES. 1973. Functional anatomy of the newborn:<br />

an atlas. New Haven: Yale <strong>University</strong> Press.


D.E. Lieberman et al.]<br />

PRIMATE CRANIAL BASE 165<br />

Crompton AW, German RZ, <strong>The</strong>xton AJ. 1997. Mechanisms<br />

of swallowing <strong>and</strong> airway protection in infant<br />

mammals (Sus domesticus <strong>and</strong> Macaca fascicularis).<br />

J Zool Lond 241:89–102.<br />

Dabelow A. 1929. Über Korrelationenin der phylogenetischen<br />

Entwicklung der Schädelform. Gegenbaurs<br />

Morphol Jahrb 63:1–49.<br />

Dabelow A. 1931. Über Korrelationenin der phylogenetischen<br />

Entwicklung der Schädelform. II. Die Beziehungen<br />

zwischen Gehirn und Schädelbasisform bei<br />

den Mammaliern. Gegenbaurs Morphol Jahrb 67:84–<br />

133.<br />

Dagosto M. 1988. Implications of post<strong>cranial</strong> evidence<br />

for the origin of eu<strong>primate</strong>s. J Hum Evol 17:35–56.<br />

David DJ, Hemmy DC, Cooter RD. 1989. Craniofacial<br />

deformities. New York: Springer Verlag.<br />

Dean MC. 1988. Growth processes in the <strong>cranial</strong> <strong>base</strong> of<br />

hominoids <strong>and</strong> their bearing on the morphological<br />

similarities that exist in the <strong>cranial</strong> <strong>base</strong> of Homo <strong>and</strong><br />

Paranthropus. In: Grine FE, editor. Evolutionary history<br />

of the “robust” Australopithecines. New York:<br />

Aldine. p 107–112.<br />

Dean MC, Wood BA. 1981. Metrical analysis of the<br />

basicranium of extant hominoids <strong>and</strong> Australopithecus.<br />

Am J Phys Anthropol 59:53–71.<br />

Dean MC, Wood BA. 1982. Basi<strong>cranial</strong> anatomy of Plio-<br />

Pleistocene hominids from East <strong>and</strong> South Africa.<br />

Am J Phys Anthropol 59:157–174.<br />

de Beer GR. 1937. <strong>The</strong> development of the vertebrate<br />

skull. Oxford: Oxford <strong>University</strong> Press.<br />

Delattre A, Fenart R. 1956. Le développement du crâne<br />

du gorille et du chimpanzé comparé au développement<br />

du crâne humain. Bull Mem Soc Anthropol<br />

Paris 6:159–173.<br />

Delattre A, Fenart R. 1963. Études des projections horizontale<br />

et vertico-frontales du crâne au cours de<br />

l’hominisation. Anthropologie 67:525–561.<br />

Demes AB. 1985. Biomechanics of the <strong>primate</strong> skull<br />

<strong>base</strong>. Advances in anatomy, embryology <strong>and</strong> cell biology.<br />

Volume 94. Berlin: Springer-Verlag.<br />

Diewert VM. 1985. Development of human craniofacial<br />

morphology during the late embryonic <strong>and</strong> early fetal<br />

periods. Am J Orthod 88:64–76.<br />

Dmoch R. 1975. Beiträge zum formenw<strong>and</strong>el des <strong>primate</strong>ncraniums<br />

mit bemerkungen zu den sagittalen<br />

knickungsverhältnissen. II Streckenanalyse, I, Neurocranium.<br />

Gegenbaurs Morphol Jahrb 121:521–601.<br />

DuBrul EL. 1950. Posture, locomotion <strong>and</strong> the skull in<br />

Lagomorpha. Am J Anat 87:277–313.<br />

DuBrul EL. 1977. Early hominid feeding mechanisms.<br />

Am J Phys Anthropol 47:305–320.<br />

DuBrul EL. 1979. Origin <strong>and</strong> adaptations of the hominid<br />

jaw joint. In: Sarnat BG, Laskin DM, editors. <strong>The</strong><br />

temporom<strong>and</strong>ibular jaw joint. 3rd ed. Springfield, IL:<br />

Thomas. p 5–34.<br />

DuBrul EL, Laskin DM. 1961. Preadaptive potentialities<br />

of the mammalian skull: an experiment in growth<br />

<strong>and</strong> form. Am J Anat 109:117–132.<br />

Duckworth WLH. 1904. Studies from the anthropological<br />

laboratory. Cambridge: Cambridge <strong>University</strong><br />

Press.<br />

Duckworth WLH. 1915. Morphology <strong>and</strong> anthropology.<br />

2nd ed. Cambridge: Cambridge <strong>University</strong> Press.<br />

Duterloo HS, Enlow DH. 1970. A comparative study of<br />

<strong>cranial</strong> growth in Homo <strong>and</strong> Macaca. Am J Anat<br />

127:357–368.<br />

Enlow DH. 1976. Postnatal growth <strong>and</strong> development of<br />

the face <strong>and</strong> cranium. In: Cohen B, Kramer IRV,<br />

editors. Scientific foundations of dentistry. London:<br />

Heinemann. p 29–46.<br />

Enlow DH. 1990. Facial growth. 3rd ed. Philadelphia:<br />

Saunders.<br />

Enlow DH, Azuma M. 1975. Functional growth boundaries<br />

in the human <strong>and</strong> mammalian face. In: Bergsma<br />

D, editor. Morphogenesis <strong>and</strong> malformation of face<br />

<strong>and</strong> brain. New York: A.R. Liss. p 217–230.<br />

Enlow DH, Bhatt MK. 1984. Facial morphology variations<br />

associated with head form variations. J Charles<br />

Tweed Found 12:21–23.<br />

Enlow DH, Hans MG. 1996. Essentials of facial growth.<br />

Philadelphia: W.B. Saunders.<br />

Fant G. 1960. Acoustic theory of speech production. <strong>The</strong><br />

Hague: Mouton.<br />

Felsenstein J. 1985. Phylogenies <strong>and</strong> the comparative<br />

method. Am Nat 125:1–15.<br />

Fenart R. 1966. Changements morphologiques de<br />

l’encéphale, chez le rat amputé des membres antérieurs.<br />

J Hirnforsch 8:493–501.<br />

Fenart R, Pellerin C. 1988. <strong>The</strong> vestibular orientation<br />

method: its application in the determination of an<br />

average human skull type. Int J Anthropol 3:223–<br />

239.<br />

Fleagle JG. 1999. Primate adaptation <strong>and</strong> evolution.<br />

2nd ed. New York: Academic Press.<br />

Ford EH. 1956. <strong>The</strong> growth of the foetal skull. J Anat<br />

90:63–72.<br />

Ford EH. 1958. Growth of the human <strong>cranial</strong> <strong>base</strong>. Am J<br />

Orthod 44:498–506.<br />

George SL. 1978. A longitudinal <strong>and</strong> cross-sectional<br />

analysis of the growth of the postnatal <strong>cranial</strong> <strong>base</strong><br />

angle. Am J Phys Anthropol 49:171–178.<br />

German RZ, Crompton AW, McCluskey C, <strong>The</strong>xton AJ.<br />

1996. Coordination between respiration <strong>and</strong> deglutition<br />

in a preterm infant mammal, Sus scrofa. Arch<br />

Oral Biol 41:619–622.<br />

Giles WB, Philips CL, Joondeph DR. 1981. Growth in<br />

the basi<strong>cranial</strong> synchondroses of adolescent Macaca<br />

mullata. Anat Rec 19:259–266.<br />

Gould SJ. 1975. Allometry in <strong>primate</strong>s, with emphasis<br />

on scaling <strong>and</strong> the evolution of the brain. Contrib<br />

Primatol 5:244–292.<br />

Gould SJ. 1977. Ontogeny <strong>and</strong> phylogeny. Cambridge:<br />

Belknap Press.<br />

Graf W. 1988. Motion detection in physical space <strong>and</strong> its<br />

peripheral <strong>and</strong> central representation. Ann NY Acad<br />

Sci 545:154–169.<br />

Graf W, deWaele C, Vidal PP, Wang DH, Evinger C.<br />

1995. <strong>The</strong> orientation of the cervical vertebral column<br />

in unrestrained awake animals. II. Movement strategies.<br />

Brain Behav Evol 45:209–231.<br />

Hall BK, editor. 1994. Homology: the hierarchical basis<br />

of comparative biology. New York: Academic Press.<br />

Hamrick MW. 1998. Functional <strong>and</strong> adaptive significance<br />

of <strong>primate</strong> pads <strong>and</strong> claws: evidence from New<br />

World anthropoids. Am J Phys Anthropol 106:113–<br />

127.<br />

Hamrick MW. 1999. Pattern <strong>and</strong> process in the evolution<br />

of <strong>primate</strong> nails <strong>and</strong> claws. J Hum Evol 37:293–<br />

297.<br />

Heintz N. 1966. Le crâne des anthropomorphes: croissance<br />

relative, variabilité, évolution. Musée Royale de<br />

l’Afrique Centrale-Tervuren. Ann Sci Zool 6:1–122.<br />

Herring SW. 1993. Epigenetic <strong>and</strong> <strong>function</strong>al influences<br />

on skull growth. In: Hanken J, Hall BK, editors. <strong>The</strong><br />

skull. Volume 1. Chicago: <strong>University</strong> of Chicago<br />

Press. p 153–206.<br />

Hofer H. 1960. Studien zum Problem des Gestaltw<strong>and</strong>els<br />

des Schädels der Säugetiere, insbesondere der<br />

Primaten. I. Die medianen Kruemmundgen des<br />

Schädels und ihr Ehrfassung nach L<strong>and</strong>zert. Z Morphol<br />

Anthropol 50:299–316.<br />

Hofer H. 1965. Die morphologische Analyse des<br />

Schädels des Menschen. In: Hereberer G, editor.


166 YEARBOOK OF PHYSICAL ANTHROPOLOGY [Vol. 43, 2000<br />

Menschliche Abstammungslehre. Stuttgart: G. Fischer.<br />

p 145–226.<br />

Hofer H. 1969. On the evolution of the craniocerebral<br />

topography in <strong>primate</strong>s. Ann NY Acad Sci 162:341–<br />

356.<br />

Hofer H, Spatz W. 1963. Studien zum Problem des<br />

Gestaltw<strong>and</strong>els des Schädels der Säugetiere, insbesondere<br />

der Primaten. II. Über die kyphosen fetaler<br />

und neonater Primatenschädel. Z Morphol Anthropol<br />

53:29–52.<br />

Hooton EA. 1946. Up from the ape. 3rd ed. New York:<br />

Macmillan.<br />

Howells WW. 1973. Cranial variation in man. Cambridge:<br />

Peabody Museum Papers, no. 67.<br />

Hoyte DA. 1991. <strong>The</strong> <strong>cranial</strong> <strong>base</strong> in normal <strong>and</strong> abnormal<br />

growth. Neurosurg Clin North Am 2:515–537.<br />

Huxley TH. 1867. On two widely contrasted forms of the<br />

human cranium. J Anat Physiol 1:60–77.<br />

Hyl<strong>and</strong>er WL, Ravosa MJ. 1992. An analysis of the<br />

supraorbital region of <strong>primate</strong>s: a morphometric <strong>and</strong><br />

experimental approach. In: Smith P, Tchernov E, editors.<br />

Structure, <strong>function</strong> <strong>and</strong> evolution of teeth. London:<br />

Freund. p 223–255.<br />

Hyl<strong>and</strong>er WL, Ravosa MJ, Ross CF, Johnson KR. 1998.<br />

M<strong>and</strong>ibular corpus strain in <strong>primate</strong>s: further evidence<br />

for a <strong>function</strong>al link between symphyseal fusion<br />

<strong>and</strong> jaw-adductor muscle force. Am J Phys Anthropol<br />

107:257–271.<br />

Hyl<strong>and</strong>er WL, Ravosa MJ, Ross CF, Wall CE, Johnson<br />

KR. 2000. Symphyseal fusion <strong>and</strong> jaw-adductor muscle<br />

force: an EMG study. Am J Phys Anthropol 112:<br />

469–492.<br />

Jeffery N. 1999. Fetal development <strong>and</strong> evolution of the<br />

human <strong>cranial</strong> <strong>base</strong>. Ph.D. dissertation, <strong>University</strong><br />

College, London.<br />

Jerison HJ. 1982. Allometry, brain size <strong>and</strong> convolutedness.<br />

In: Armstrong E, Falk D, editors. Primate brain<br />

evolution. Methods <strong>and</strong> concepts. New York: Plenum<br />

Press. p 77–84.<br />

Jolicoeur P, Mossiman J. 1968. Intervalles de confidance<br />

pour la peute de l’axe majeur d’une distribution<br />

normale bidimensionelle. Biometrie Praximetrie<br />

9:121–140.<br />

K<strong>and</strong>el ER, Schwartz JH, Jessell TM. 1991. Principles<br />

of neural science. 3rd ed. Norwalk, CT: Apple <strong>and</strong><br />

Lang.<br />

Kay RF, Cartmill M. 1974. Skull of Palaechthon nacimienti.<br />

Nature 252:37–38.<br />

Kay RF, Cartmill M. 1977. Cranial morphology <strong>and</strong><br />

adaptations of Palaechthon nacimienti <strong>and</strong> other<br />

Paromomyidae (Plesiadapoidea, ?Primates), with a<br />

description of a new genus <strong>and</strong> species. J Hum Evol<br />

6:19–35.<br />

Kjaer I. 1990. Ossification of the human basicranium. J<br />

Craniofac Genet Dev Biol 10:29–38.<br />

Kodama G. 1976a. Developmental studies of the presphenoid<br />

of the human sphenoid bone. In: Bosma JF,<br />

editor. Development of the basicranium. Bethesda,<br />

MD: National Institutes of Health. p 141–155.<br />

Kodama G. 1976b. Developmental studies of the body of<br />

the human sphenoid bone. In: Bosma JF, editor. Development<br />

of the basicranium. Bethesda, MD: National<br />

Institutes of Health. p. 156–165.<br />

Kodama G. 1976c. Developmental studies of the orbitosphenoid<br />

of the human sphenoid bone. In: Bosma JF,<br />

editor. Development of the basicranium. Bethesda,<br />

MD: National Institutes of Health. p 166–176.<br />

Kohn LAP, Leigh SR, Jacobs SC, Cheverud JM. 1993.<br />

Effects of annular <strong>cranial</strong> vault modification on the<br />

<strong>cranial</strong> <strong>base</strong> <strong>and</strong> face. Am J Phys Anthropol 90:147–<br />

168.<br />

Krogman WM. 1969. Growth changes in the skull, face,<br />

jaws <strong>and</strong> teeth of the chimpanzee. In: Bourne GH,<br />

editor. <strong>The</strong> chimpanzee. Volume 1. Basel: S. Karger. p<br />

104–164.<br />

Lager H. 1958. A histological description of the <strong>cranial</strong><br />

<strong>base</strong> in Macaca rhesus. Eur Orthod Soc Trans 34:147–<br />

156.<br />

Laitman JT, Heimbuch RC. 1982. <strong>The</strong> basicranium of<br />

Plio-Pleistocene hominids as an indicator of their upper<br />

respiratory systems. Am J Phys Anthropol 59:<br />

323–343.<br />

Laitman JT, Reidenberg JS. 1993. Specializations of the<br />

human upper respiratory <strong>and</strong> upper digestive systems<br />

as seen through comparative <strong>and</strong> developmental<br />

anatomy. Dysphagia 8:318–325.<br />

Laitman JT, Heimbuch RC, Crelin ES. 1978. Developmental<br />

change in a basi<strong>cranial</strong> line <strong>and</strong> its relationship<br />

to the upper respiratory system in living <strong>primate</strong>s.<br />

Am J Anat 152:467–483.<br />

Laitman JT, Heimbuch RC, Crelin ES. 1979. <strong>The</strong> basicranium<br />

of fossil hominids as an indicator of their<br />

upper respiratory systems. Am J Phys Anthropol 51:<br />

15–34.<br />

L<strong>and</strong>zert T. 1866. Der Sattelwinkel und sein Verhaeltnis<br />

zur Pro- und Orthognathie. Abh Senckenberg<br />

Naturforsch Ges 6:19–165.<br />

Langille RM, Hall BK. 1993. Pattern formation <strong>and</strong> the<br />

neural crest. In: Hanken J, Hall BK, editors. <strong>The</strong><br />

skull. Volume 1. Chicago: <strong>University</strong> of Chicago<br />

Press. p 77–111.<br />

Larson JE, Herring SW. 1996. Movement of the epiglottis<br />

in mammals. Am J Phys Anthropol 100:71–82.<br />

Latham RA. 1972. <strong>The</strong> sella point <strong>and</strong> postnatal growth<br />

of the human <strong>cranial</strong> <strong>base</strong>. Am J Orthod 61:156–162.<br />

Le Douarin NM, Ziller C, Couly GF. 1993. Patterning of<br />

neural crest derivatives in the avian embryo: in vivo<br />

<strong>and</strong> in vitro studies. Dev Biol 159:24–49.<br />

Lele S. 1993. Euclidean distance matrix analysis<br />

(EDMA): estimation of mean form <strong>and</strong> mean form<br />

difference. Math Geol 25:573–602.<br />

Lemelin P. 1999. Morphological correlates of substrate<br />

use in didelphid marsupials: implications for <strong>primate</strong><br />

origins. J Zool Lond 247:165–175.<br />

Lieberman DE. 1995. Testing hypotheses about recent<br />

human evolution using skulls: integrating morphology,<br />

<strong>function</strong>, development, <strong>and</strong> phylogeny. Curr Anthropol<br />

36:159–197.<br />

Lieberman DE. 1997. Making behavioral <strong>and</strong> phylogenetic<br />

inferences from fossils: considering the developmental<br />

influence of mechanical forces. Annu Rev Anthropol<br />

26:185–210.<br />

Lieberman DE. 1998. Sphenoid shortening <strong>and</strong> the evolution<br />

of modern human <strong>cranial</strong> shape. Nature 393:<br />

158–162.<br />

Lieberman DE. 1999. Homology <strong>and</strong> hominid phylogeny:<br />

problems <strong>and</strong> potential solution. Evol Anthropol<br />

7:142–151.<br />

Lieberman DE. 2000. Ontogeny, homology, <strong>and</strong> phylogeny<br />

in the hominid craniofacial skeleton: the problem<br />

of the browridge. In: O’Higgins P, Cohn M, editors.<br />

Development, growth <strong>and</strong> evolution: implications for<br />

the study of hominid skeletal evolution. London: Academic<br />

Press. p 85–122.<br />

Lieberman DE, McCarthy RC. 1999. <strong>The</strong> <strong>ontogeny</strong> of<br />

<strong>cranial</strong> <strong>base</strong> angulation in humans <strong>and</strong> chimpanzees<br />

<strong>and</strong> its implications for reconstructing pharyngeal<br />

dimensions. J Hum Evol 36:487–517.<br />

Lieberman DE, Pilbeam DR, Wood BA. 1996. Homoplasy<br />

<strong>and</strong> early Homo: an analysis of the evolutionary<br />

relationships of H. habilis sensu stricto <strong>and</strong><br />

H. rudolfensis. J Hum Evol 30:97–120.


D.E. Lieberman et al.]<br />

PRIMATE CRANIAL BASE 167<br />

Lieberman DE, McCarthy RC, Hiiemae K, Lieberman<br />

P, Palmer JB. 1998. New estimates of fossil hominid<br />

vocal tract dimension. J Hum Evol 34:12–13.<br />

Lieberman DE, Pearson OM, Mowbray KM. 2000. Basi<strong>cranial</strong><br />

influence on overall <strong>cranial</strong> shape. J Hum<br />

Evol 38:291–315.<br />

Lieberman P. 1984. <strong>The</strong> biology <strong>and</strong> evolution of language.<br />

Cambridge, MA: <strong>Harvard</strong> <strong>University</strong> Press.<br />

Lieberman P, Crelin ES. 1971. On the speech of Ne<strong>and</strong>erthal<br />

man. Ling Inq 1:203–223.<br />

Lieberman P, Crelin ES, Klatt DH. 1972. Phonetic ability<br />

<strong>and</strong> related anatomy of the newborn <strong>and</strong> adult<br />

human, Ne<strong>and</strong>erthal man, <strong>and</strong> the chimpanzee. Am<br />

Anthropol 74:287–307.<br />

Lieberman P, Laitman JT, Reidenberg JS, Gannon PJ.<br />

1992. <strong>The</strong> anatomy, physiology, acoustics <strong>and</strong> perception<br />

of speech: essential elements in analysis of the<br />

evolution of human speech. J Hum Evol 23:447–467.<br />

Lozanoff S. 1993. Midfacial retrusion in adult Brachyrrhine<br />

mice. Acta Anat (Basel) 147:125–132.<br />

Lozanoff S, Jureczek S, Feng T, Padwal R. 1994. Anterior<br />

<strong>cranial</strong> <strong>base</strong> morphology in mice with midfacial<br />

retrusion. Cleft Palate Craniofac J 31:417–428.<br />

Lufkin T, Mark M, Hart CP, Dollé P, LeMeur M, Chambon<br />

P. 1992. Homeotic transformation of the occipital<br />

bones by ectopic expression of a homeobox gene. Nature<br />

359:835–841.<br />

Lugoba SA, Wood BA. 1990. Position <strong>and</strong> orientation of<br />

the foramen magnum in higher <strong>primate</strong>s. Am J Phys<br />

Anthropol 81:67–76.<br />

Ma W, Lozanoff S. 1999. Spatial <strong>and</strong> temporal distribution<br />

of cellular proliferation in the anterior <strong>cranial</strong><br />

<strong>base</strong> of normal midfacially retrusive mice. Clin Anat<br />

12:315–325.<br />

Martin RD. 1990. Primate origins <strong>and</strong> evolution.<br />

Princeton: Princeton <strong>University</strong> Press.<br />

May R, Sheffer DB. 1999. Growth changes in measurements<br />

of upper facial positioning. Am J Phys Anthropol<br />

108:269–280.<br />

Maynard Smith J, Burian R, Kauffman S, Alberch P,<br />

Campbell J, Goodwin B, L<strong>and</strong>e R, Raup D, Wolpert L.<br />

1985. Developmental constraints <strong>and</strong> evolution. Q<br />

Rev Biol 60:265–287.<br />

McCarthy R. 2001. Anthropoid <strong>cranial</strong> <strong>base</strong> architecture<br />

<strong>and</strong> scaling relationships. J Hum Evol (in press).<br />

McCarthy RC, Lieberman DE. 2001. <strong>The</strong> posterior maxillary<br />

(PM) plane <strong>and</strong> anterior <strong>cranial</strong> architecture in<br />

<strong>primate</strong>s. Anat Rec (in press).<br />

Melsen B. 1969. Time of closure of the spheno-occipital<br />

synchondrosis determined on dry skulls. A radiographic<br />

craniometric study. Acta Odontol Sc<strong>and</strong> 27:<br />

73–90.<br />

Melsen B. 1971. <strong>The</strong> postnatal growth of the <strong>cranial</strong><br />

<strong>base</strong> in Macaca rhesus analyzed by the implant<br />

method. T<strong>and</strong>laegebladet 75:1320–1329.<br />

Michejda M. 1971. Ontogenetic changes of the <strong>cranial</strong><br />

<strong>base</strong> in Macaca mulatta. In: Biegert J, Leutenegger<br />

W, editors. Proceedings of the Third International<br />

Congress of Primatology Zurich 1970. Basel: Karger.<br />

p 215–225.<br />

Michejda M. 1972a. <strong>The</strong> role of basi<strong>cranial</strong> synchondroses<br />

in flexure processes <strong>and</strong> ontogenetic development<br />

of the skull <strong>base</strong>. Am J Phys Anthropol 37:143–150.<br />

Michejda M. 1972b. Significance of basi<strong>cranial</strong> synchondroses<br />

in nonhuman <strong>primate</strong>s <strong>and</strong> man. Proc 3rd<br />

Conf Exp Med Surg Primates, Lyon 1:372–378.<br />

Michejda M, Lamey D. 1971. Flexion <strong>and</strong> metric age<br />

changes of the <strong>cranial</strong> <strong>base</strong> in Macaca mulatta. Infants<br />

<strong>and</strong> juveniles. Folia Primatol (Basel) 34:133–<br />

141.<br />

Mitchison G. 1991. Neuronal branching patterns <strong>and</strong><br />

the economy of cortical wiring. Proc R Soc Lond [Biol]<br />

245:151–158.<br />

Moore WJ, Lavelle CJB. 1974. Growth of the facial<br />

skeleton in the Hominoidea. London: Academic Press.<br />

Moore WJ, Adams LM, Lavelle CLB. 1973. Head posture<br />

in the Hominoidea. J Zool Lond 169:409–416.<br />

Moss ML. 1958. <strong>The</strong> pathogenesis of artificial <strong>cranial</strong><br />

deformation. Am J Phys Anthropol 16:269–286.<br />

Moss ML. 1961. Rotation of the otic capsule in bipedal<br />

rats. Am J Phys Anthropol 19:301–307.<br />

Moss ML. 1963. Morphological variations of the crista<br />

galli <strong>and</strong> medial orbital margin. Am J Phys Anthropol<br />

21:259–264.<br />

Moss ML. 1976. Experimental alteration of basi-synchondroseal<br />

cartilage growth in rat <strong>and</strong> mouse. In:<br />

Bosma JF, editor. Development of the basicranium.<br />

Bethesda, MD: National Institutes of Health. p 541–<br />

569.<br />

Moss ML. 1997. <strong>The</strong> <strong>function</strong>al matrix hypothesis revisited.<br />

1. <strong>The</strong> role of mechanotransduction. Am J<br />

Orthod Dentofacial Orthop 112:8–11.<br />

Moss ML, Young RW. 1960. A <strong>function</strong>al approach to<br />

craniology. Am J Phys Anthropol 18:281–292.<br />

Moss ML, Moss-Salentjin L, Vilmann H, Newell-Morris<br />

L. 1982. Neuro-skeletal topology of the <strong>primate</strong> basicranium:<br />

its implications for the “fetalization hypothesis.”<br />

Gegenbaurs Morphol Jahrb 128:58–67.<br />

Nearey T. 1978. Phonetic features for vowels. Bloomington:<br />

Indiana <strong>University</strong> Linguistics Club.<br />

Negus V. 1949. <strong>The</strong> comparative anatomy <strong>and</strong> physiology<br />

of the larynx. New York: Hafner.<br />

Noble VE, Kowalski EM, Ravosa MJ. 2000. Orbital orientation<br />

<strong>and</strong> the <strong>function</strong> of the mammalian postorbital<br />

bar. J Zool Lond 250:405–418.<br />

Noden DM. 1991. Cell movements <strong>and</strong> control of patterned<br />

tissue assembly during craniofacial development.<br />

J Craniofac Genet Dev Biol 11:192–213.<br />

Nunn CL. 1995. A simulation test of Smith’s “degress of<br />

freedom” correction for comparative studies. Am J<br />

Phys Anthropol 98:355–367.<br />

O’Higgins P. 2000. Quantitative approaches to the<br />

study of craniofacial growth <strong>and</strong> evolution: advances<br />

in morphometric techniques. In: O’Higgins P, Cohn<br />

M, editors. Development, growth <strong>and</strong> evolution: implications<br />

for the study of hominid skeletal evolution.<br />

London: Academic Press. p 163–185.<br />

Olson E, Miller R. 1958. Morphological integration. Chicago:<br />

<strong>University</strong> of Chicago Press.<br />

Olson TR. 1985. Cranial morphology <strong>and</strong> systematics of<br />

the Hadar hominids <strong>and</strong> “Australopithecus” africanus.<br />

In: Delson E, editor. Ancestors: the hard evidence.<br />

New York: A.R. Liss. p 102–119.<br />

O’Rahilly R, Müller F. 1999. <strong>The</strong> embryonic human<br />

brain: an atlas of developmental stages. 2nd ed. New<br />

York: Wiley-Liss.<br />

Purvis A, Rambaut A. 1995. Comparative analysis by<br />

independent contrasts (CAIC): an Apple Macintosh<br />

application for analyzing comparative data. Comp<br />

Appl Biosci 11:247–251.<br />

Radinsky LB. 1968. A new approach to mammalian<br />

<strong>cranial</strong> analysis, illustrated by examples of prosimian<br />

<strong>primate</strong>s. J Morphol 124:167–179.<br />

Radinsky LB. 1970. <strong>The</strong> fossil record of prosimian brain<br />

evolution. In: Noback C, Montagna W, editors. <strong>The</strong><br />

<strong>primate</strong> brain. Advances in primatology. Volume 1.<br />

New York: Appleton-Century-Crofts. p 209–224.<br />

Radinsky LB. 1977. Early <strong>primate</strong> brains: facts <strong>and</strong><br />

fiction. J Hum Evol 6:79–86.<br />

Radinsky LB. 1979. <strong>The</strong> fossil evidence of <strong>primate</strong> brain<br />

evolution (James Arthur Lecture). New York: American<br />

Museum of Natural History.


168 YEARBOOK OF PHYSICAL ANTHROPOLOGY [Vol. 43, 2000<br />

Ranke J. 1892. Über einige gesetzmassige Beziehungen<br />

zwischen Schädelgrund, Gehirn- und Gesichtsschädel.<br />

Beitrage Anthropol Urgesch Bayerns 12:1–132.<br />

Ravosa MJ. 1988. Browridge development in Cercopithecidae:<br />

a test of two models. Am J Phys Anthropol<br />

76:535—555.<br />

Ravosa MJ. 1991a. Ontogenetic perspective on mechanical<br />

<strong>and</strong> nonmechanical models of <strong>primate</strong> circumorbital<br />

morphology. Am J Phys Anthropol 85:95–112.<br />

Ravosa MJ. 1991b. Interspecific perspective on mechanical<br />

<strong>and</strong> nonmechanical models of <strong>primate</strong> circumorbital<br />

morphology. Am J Phys Anthropol 86:369–396.<br />

Ravosa MJ. 1996. M<strong>and</strong>ibular form <strong>and</strong> <strong>function</strong> in<br />

North American <strong>and</strong> European Adapidae <strong>and</strong> Omomyidae.<br />

J Morphol 229:171–190.<br />

Ravosa MJ. 1999. Anthropoid origins <strong>and</strong> the modern<br />

symphysis. Folia Primatol (Basel) 70:65–78.<br />

Ravosa MJ, Shea BT. 1994. Pattern in craniofacial biology:<br />

evidence from the Old World monkeys (Cercopithecidae).<br />

Int J Primatol 15:801–822.<br />

Ravosa MJ, Noble VE, Hyl<strong>and</strong>er WL, Johnson KR, Kowalski<br />

EM. 2000a. Masticatory stress, orbital orientation<br />

<strong>and</strong> the evolution of the <strong>primate</strong> postorbital<br />

bar. J Hum Evol 38:667–693.<br />

Ravosa MJ, Vinyard CJ, Hyl<strong>and</strong>er WL. 2000b. Stressed<br />

out: masticatory forces <strong>and</strong> <strong>primate</strong> circumorbital<br />

form. Anat Rec (New Anat) 261:(in press).<br />

Reidenberg JS, Laitman JT. 1991. Effect of basi<strong>cranial</strong><br />

flexion on larynx <strong>and</strong> hyoid position in rats: an experimental<br />

study of skull <strong>and</strong> soft tissue interactions.<br />

Anat Rec 230:557–569.<br />

Riesenfeld A. 1966. <strong>The</strong> effects of experimental bipedalism<br />

<strong>and</strong> upright posture in the rat <strong>and</strong> their significance<br />

for human evolution. Acta Anat (Basel) 65:449–521.<br />

Riesenfeld A. 1969. <strong>The</strong> adaptive m<strong>and</strong>ible: an experimental<br />

study. Acta Anat (Basel) 72:246–262.<br />

Riolo ML, Moyers RE, McNamara JA Jr, Hunter WS.<br />

1974. An atlas of craniofacial growth: cephalometric<br />

st<strong>and</strong>ards from the <strong>University</strong> School Growth Study,<br />

the <strong>University</strong> of Michigan. Ann Arbor: Center for<br />

Human Growth <strong>and</strong> Development, monograph no. 2.<br />

Robinson GW, Mahon KA. 1994. Differential <strong>and</strong> overlapping<br />

expression domains of DLx2 <strong>and</strong> Dlx3 suggest<br />

distinct roles for distal-less genes in craniofacial development.<br />

Mech Dev 48:199–215.<br />

Rosenberger AL. 1985. In favor of the Necrolemur-<br />

Tarsier hypothesis. Folia Primatol (Basel) 45:179–194.<br />

Rosenberger AL. 1986. Platyrrhines, catarrhines <strong>and</strong><br />

the anthropoid transition. In: Wood BA, Martin L,<br />

Andrews P, editors. Major topics in <strong>primate</strong> <strong>and</strong> human<br />

evolution. Cambridge: Cambridge <strong>University</strong><br />

Press. p 66–88.<br />

Ross CF. 1994. Craniofacial evidence for anthropoid <strong>and</strong><br />

tarsier relationships. In: Fleagle JG, Kay RF, editors.<br />

Anthropoid origins. New York: Plenum Press. p 469–<br />

547.<br />

Ross CF. 1995a. Allometric <strong>and</strong> <strong>function</strong>al influences on<br />

<strong>primate</strong> orbit orientation <strong>and</strong> the origins of the Anthropoidea.<br />

J Hum Evol 29:201–227.<br />

Ross CF. 1995b. Muscular <strong>and</strong> osseous anatomy of the<br />

<strong>primate</strong> anterior temporal fossa <strong>and</strong> the <strong>function</strong>s of<br />

the postorbital septum. Am J Phys Anthropol 98:275–<br />

306.<br />

Ross CF. 1996. An adaptive explanation for the origin of<br />

the Anthropoidea (Primates). Am J Primatol 40:205–<br />

230.<br />

Ross CF. 2000. Into the light: the origin of Anthropoidea.<br />

Annu Rev Anthropol 29:147–194.<br />

Ross CF, Henneberg M. 1995. Basi<strong>cranial</strong> flexion, relative<br />

brain size <strong>and</strong> facial kyphosis in Homo sapiens<br />

<strong>and</strong> some fossil hominids. Am J Phys Anthropol 98:<br />

575–593.<br />

Ross CF, Ravosa MJ. 1993. Basi<strong>cranial</strong> flexion, relative<br />

brain size <strong>and</strong> facial kyphosis in nonhuman <strong>primate</strong>s.<br />

Am J Phys Anthropol 91:305–324.<br />

Ross CF, Williams BA, Kay RF. 1998. Phylogenetic<br />

analysis of anthropoid relationships. J Hum Evol 35:<br />

221–306.<br />

Ruff CB, Trinkaus E, Holliday TW. 1997. Body mass<br />

<strong>and</strong> encephalization in Pleistocene Homo. Nature<br />

387:173–176.<br />

Sakka M, Aaron C, Hervé D. 1976. Étude anatomique,<br />

fonctionelle et radiologique de l’ensemble cervicocéphalique<br />

dorsal du Chien. Applications aux problèmes<br />

du port de tête en station érigrée chez les Hominidés.<br />

Zbl Vet Med C 5:1420.<br />

Sarnat BG. 1982. Eye <strong>and</strong> orbital size in the young <strong>and</strong><br />

adult. Some postnatal experimental <strong>and</strong> clinical relationships.<br />

Ophthalmology 185:74–89.<br />

Sarnat BG. 1988. Craniofacial change <strong>and</strong> non-change<br />

after experimental surgery in young <strong>and</strong> adult animals.<br />

Angle Orthod 58:321–342.<br />

Sarnat BG. 1999. Basic science <strong>and</strong> clinical experimental<br />

<strong>primate</strong> studies in craniofaciodental biology: a<br />

personal historical review. J Oral Maxillofac Surg<br />

57:714–724.<br />

Sasaki H, Kodama G. 1976. Developmental studies of<br />

the postsphenoid of the human sphenoid bone. In:<br />

Bosma JF, editor. Development of the basicranium.<br />

Bethesda, MD: National Institutes of Health. p. 177–<br />

190.<br />

Schilling TF, Thorogood PV. 2000. Development <strong>and</strong><br />

evolution of the vertebrate skull. In: O’Higgins P,<br />

Cohn M, editors. Development, growth <strong>and</strong> evolution:<br />

implications for the study of hominid skeletal evolution.<br />

London: Academic Press. p 57–83.<br />

Schneiderman ED. 1992. Facial growth in the rhesus<br />

monkey: a longitudinal <strong>and</strong> cephalometric study.<br />

Princeton: Princeton <strong>University</strong> Press.<br />

Schultz AH. 1940. <strong>The</strong> size of the orbit <strong>and</strong> of the eye in<br />

<strong>primate</strong>s. Am J Phys Anthropol 26:398–408.<br />

Schultz AH. 1942. Conditions for balancing the head in<br />

<strong>primate</strong>s. Am J Phys Anthropol 29:483–497.<br />

Schultz AH. 1955. <strong>The</strong> position of the occipital condyles<br />

<strong>and</strong> of the face relative to the skull <strong>base</strong> in <strong>primate</strong>s.<br />

Am J Phys Anthropol 13:97–120.<br />

Scott JH. 1958. <strong>The</strong> <strong>cranial</strong> <strong>base</strong>. Am J Phys Anthropol<br />

16:319–348.<br />

Shapiro R. 1960. <strong>The</strong> normal skull: a Roentgen study.<br />

New York: Hoeber.<br />

Shapiro R, Robinson F. 1980. <strong>The</strong> embryogenesis of the<br />

human skull. Cambridge: <strong>Harvard</strong> <strong>University</strong> Press.<br />

Shea BT. 1985a. On aspects of skull form in African<br />

apes <strong>and</strong> orangutans, with implications for hominoid<br />

evolution. Am J Phys Anthropol 68:329–342.<br />

Shea BT. 1985b. Ontogenetic allometry <strong>and</strong> scaling: a<br />

discussion <strong>base</strong>d on the growth <strong>and</strong> form of the skull<br />

in African apes. In: Jungers WL, editor. Size <strong>and</strong><br />

scaling in <strong>primate</strong> biology. New York: Plenum Press.<br />

p 175–205.<br />

Shea BT. 1986. On skull form <strong>and</strong> the supraorbital<br />

torus in <strong>primate</strong>s. Curr Anthropol 27:257–259.<br />

Shea BT. 1987. Reproductive strategies, body size, <strong>and</strong><br />

encephalization in <strong>primate</strong> evolution. Int J Primatol<br />

8:139–156.<br />

Shea BT. 1988. Phylogeny <strong>and</strong> skull form in the hominoid<br />

<strong>primate</strong>s. In: Schwartz JH, editor. <strong>The</strong> biology of<br />

the orangutan. New York: Oxford <strong>University</strong> Press. p<br />

233–246.<br />

Simons EL. 1962. Fossil evidence relating to the early<br />

evolution of <strong>primate</strong> behavior. Ann NY Acad Sci 102:<br />

282–294.<br />

Sirianni J. 1985. Nonhuman <strong>primate</strong>s as models for<br />

human craniofacial growth. In: Watt ES, editor. Non-


D.E. Lieberman et al.]<br />

PRIMATE CRANIAL BASE 169<br />

human <strong>primate</strong> models for human growth <strong>and</strong> development.<br />

New York: A.R. Liss. p 95–124.<br />

Sirianni J, Newell-Morris L. 1980. Craniofacial growth<br />

of fetal Macaca nemestrina: a cephalometric roentgenographic<br />

study. Am J Phys Anthropol 53:407–421.<br />

Sirianni JE, Swindler DR. 1979. A review of postnatal<br />

craniofacial growth in Old World monkeys <strong>and</strong> apes.<br />

Yrbk Phys Anthropol 22:80–104.<br />

Sirianni JE, Swindler DR. 1985. Growth <strong>and</strong> development<br />

of the pigtailed macaque. Boca Raton: CRC Press.<br />

Sirianni JE, Van Ness AL. 1978. Postnatal growth of<br />

the <strong>cranial</strong> <strong>base</strong> in Macaca nemestrina. Am J Phys<br />

Anthropol 49:329–340.<br />

Sjøvold T. 1984. A report on the heritability of some<br />

<strong>cranial</strong> measurements <strong>and</strong> non-metric traits. In: Van<br />

Vark GN, Howells WW, editors. Multivariate statistics<br />

in physical anthropology. Dordrecht: D. Reidel. p<br />

223–246.<br />

Slavkin HC. 1989. Developmental craniofacial biology.<br />

Philadelphia: Lea <strong>and</strong> Febiger.<br />

Smith KK. 1996. Integration of craniofacial structures<br />

during development in mammals. Am Zool 26:70–79.<br />

Smith RJ. 1994. Degrees of freedom in interspecific<br />

allometry: an adjustment for the effects of phylogenetic<br />

constraint. Am J Phys Anthropol 93:95–107.<br />

Smith RJ, Josell SD. 1984. <strong>The</strong> plan of the human face: a<br />

test of three general concepts. Am J Orthod 85:103–108.<br />

Sokal RF, Rohlf FJ. 1981. Biometry. 2nd ed. New York:<br />

W.H. Freeman <strong>and</strong> Co.<br />

Sperber GH. 1989. Craniofacial embryology. 4th ed.<br />

London: Wright.<br />

Spoor CF. 1993. <strong>The</strong> comparative morphology <strong>and</strong> phylogeny<br />

of the human bony labyrinth. Ph.D. dissertation,<br />

<strong>University</strong> of Utrecht.<br />

Spoor CF. 1997. Basi<strong>cranial</strong> architecture <strong>and</strong> relative<br />

brain size of Sts 5 (Australopithecus africanus) <strong>and</strong><br />

other Plio-Pleistocene hominids. S Afr J Sci 93:182–186.<br />

Spoor CF, Zonneveld F. 1998. A comparative review of<br />

the human bony labyrinth. Yrbk Phys Anthropol 41:<br />

211–251.<br />

Spoor CF, O’Higgins P, Dean MC, Lieberman DE. 1999.<br />

Anterior sphenoid in modern humans. Nature 397:572.<br />

Spoor CF, Jeffery N, Zonneveld F. 2000. Imaging skeletal<br />

growth <strong>and</strong> evolution. In: O’Higgins P, Cohn M,<br />

editors. Development, growth <strong>and</strong> evolution: implications<br />

for the study of hominid skeletal evolution. London:<br />

Academic Press. p 123–161.<br />

Srivastava HC. 1992. Ossification of the membranous<br />

portion of the squamous part of the occipital bone in<br />

man. J Anat 180:219–224.<br />

Stamrud L. 1959. External <strong>and</strong> internal <strong>cranial</strong> <strong>base</strong>.<br />

Acta Odontol Sc<strong>and</strong> 17:239–266.<br />

Starck D. 1975. <strong>The</strong> skull of the fetal chimpanzee. In:<br />

Bourne GH, editor. <strong>The</strong> chimpanzee. Volume 6. Basel:<br />

S. Karger. p 1–33.<br />

Stephan H, Frahm H, Baron G. 1981. New <strong>and</strong> revised<br />

data on volumes of brain structure in insectivores <strong>and</strong><br />

<strong>primate</strong>s. Comp Primate Biol 4:1–38.<br />

Stevens KN. 1972. Quantal nature of speech. In: David<br />

EE Jr, Denes PB, editors. Human communication: a<br />

unified view. New York: McGraw Hill.<br />

Stevens KN, House AS. 1955. Development of a quantitative<br />

description of vowel articulation. J Acoust Soc<br />

Am 27:484–493.<br />

Strait DS. 1998. Evolutionary integration in the hominid<br />

<strong>cranial</strong> <strong>base</strong>. Ph.D. dissertation, State <strong>University</strong><br />

of New York, Stony Brook.<br />

Strait DS. 1999. <strong>The</strong> scaling of basi<strong>cranial</strong> flexion <strong>and</strong><br />

length. J Hum Evol 37:701–719.<br />

Strait DS, Ross CF. 1999. Kinematic data on <strong>primate</strong><br />

head <strong>and</strong> neck posture: implications for the evolution<br />

of basi<strong>cranial</strong> flexion <strong>and</strong> an evaluation of registration<br />

planes. Am J Phys Anthropol 108:205–222.<br />

Strait DS, Grine FE, Moniz MA. 1997. A reappraisal of<br />

hominid phylogeny. J Hum Evol 32:17–82.<br />

Szalay FS, Rosenberger AL, Dagosto M. 1987. Diagnosis<br />

<strong>and</strong> differentiation of the order Primates. Ybk<br />

Phys Anthropol 30:75–105.<br />

Topinard P. 1890. Anthropology. London: Chapman <strong>and</strong><br />

Hall.<br />

van den Eynde B, Kjaer I, Solo B, Graem N, Kjaer TW,<br />

Mathiesen M. 1992. Cranial <strong>base</strong> angulation <strong>and</strong><br />

prognathism related to <strong>cranial</strong> <strong>and</strong> general skeletal<br />

maturation in human fetuses. J Craniofac Genet Dev<br />

Biol 12:22–32.<br />

Van der Linden FPG, Enlow DH. 1971. A study of the<br />

anterior <strong>cranial</strong> <strong>base</strong>. Am J Orthod 41:119–124.<br />

Van Essen DC. 1997. A tension-<strong>base</strong>d theory of morphogenesis<br />

<strong>and</strong> compact wiring in the central nervous<br />

system. Nature 385:313–318.<br />

Vidal PP, Graf W, Berthoz A. 1986. <strong>The</strong> orientation of the<br />

cervical vertebral column in unrestrained awake animals.<br />

I. Resting position. Exp Brain Res 61:549–559.<br />

Vielle-Grosjean I, Hunt P, Gulisano M, Boncinelli E,<br />

Thorogood PV. 1997. Branchial HOX gene expression<br />

<strong>and</strong> human craniofacial development. Dev Biol 183:<br />

49–60.<br />

Vinyard CI. 1994. A quantitative assessment of the<br />

supraorbital region in moden Melanesians. MA <strong>The</strong>sis<br />

Anthropology, Northern Illinois <strong>University</strong>.<br />

Vinyard CJ, Smith FH. 1997. Morphometric relationships<br />

between the supraorbital region <strong>and</strong> frontal<br />

sinus in Melanesian crania. Homo 48:1–21.<br />

Virchow R. 1857. Untersuchungen über die Entwicklung<br />

des Schädelgrundes im gesunden und<br />

Frankhaften Zust<strong>and</strong>e. Berlin: Gesichissiblung und<br />

Gehirnbru.<br />

Vogel C. 1964. Über eine Schädelbasisanomalie bei<br />

einem freier Wildbahn geschossen Cercopithecus torquatus<br />

atys. Z Morphol Anthropol 55:262–276.<br />

Wagner GP. 1989. <strong>The</strong> biological homology concept.<br />

Annu Rev Ecol Syst 20:51–69.<br />

Weidenreich F. 1941. <strong>The</strong> brain <strong>and</strong> its role in the<br />

phylogenetic transformation of the human skull.<br />

Trans Am Philos Soc 31:321–442.<br />

White TD, Suwa G, Berhane A. 1994. Australopithecus<br />

ramidus, a new species of early hominid from Aramis,<br />

Ethiopia. Nature 371:306–312.<br />

Williams PL, Bannister LH, Berry MM, Collins P,<br />

Dyson M, Dussek JE, Ferguson MWJ. 1995. Gray’s<br />

anatomy. 38th ed. Edinburgh: Churchill Livingstone.<br />

Young RW. 1959. <strong>The</strong> influence of <strong>cranial</strong> contents on<br />

postnatal growth of the skull in the rat. Am J Anat<br />

105:383–390.<br />

Zelditch ML. 1987. Evaluating models of developmental<br />

integration in the laboratory rat using confirmatory<br />

factor analysis. Syst Zool 36:368–380.<br />

Zelditch ML. 1988. Ontogenetic variation in patterns of<br />

phenotypic integration in the laboratory rat. Evolution<br />

42:28–41.<br />

Zelditch ML, Fink WL. 1995. Allometry <strong>and</strong> developmental<br />

integration of body growth in a pirhana, Pygocentrus<br />

natterieri (Teleostei: Ostariophysi). J Morphol<br />

223:341–355.<br />

Zelditch ML, Fink WL. 1996. Heterochrony <strong>and</strong> heterotopy:<br />

stability <strong>and</strong> innovation in the evolution of<br />

form. Paleobiology 22:241–254.<br />

Zuckerman S. 1954. Correlation of change in the evolution<br />

of higher <strong>primate</strong>s. In: Huxley JS, Hardy A, Ford<br />

EB, editors. Evolution as process. London: Allen <strong>and</strong><br />

Unwin. p 347–401.<br />

Zuckerman S. 1955. Age changes in the basi<strong>cranial</strong> axis<br />

of the human skull. Am J Phys Anthropol 13:521–539.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!