<<

Supplementary Information for

An Abundance of Developmental Anomalies and Abnormalities in Pleistocene People Erik Trinkaus Department of Anthropology, Washington University, Saint Louis MO 63130

Corresponding author: Erik Trinkaus Email: [email protected]

This PDF file includes:

Supplementary text Figures S1 to S57 Table S1 References 1 to 421 for SI reference citations

Introduction Although they have been considered to be an inconvenience for the morphological analysis of paleontological remains, it has become appreciated that various pathological lesions and other abnormalities or rare variants in human fossil remains might provide insights into Pleistocene human biology and behavior (following similar trends in bioarcheology). In this context, even though there were earlier paleopathological assessments in monographic treatments of human remains (e.g., 1-3), it has become common to provide details on abnormalities in primary descriptions of human fossils (e.g., 4-12), as well as assessments of specific lesions on known and novel remains [see references in Wu et al. (13, 14) and below]. These works have been joined by doctoral dissertation assessments of patterns of Pleistocene human lesions (e.g., 15-18). The paleopathological attention has been primarily on the documentation and differential diagnosis of the abnormalities of individual fossil remains, leading to the growing paleopathological literature on Pleistocene specimens and their lesions. There have been some considerations of the overall patterns of the lesions, but those assessments have been concerned primarily with non-specific stress indicators and traumatic lesions (e.g., 13, 15, 19-21), with variable considerations of issues of survival

1 w ww.pnas.org/cgi/doi/10.1073/pnas.1814989115 and especially the inferred social support of the afflicted (e.g., 22-27). At the same time, it has been noted that there appears to be an elevated level of abnormalities among Pleistocene , both ones that are clearly pathological and others that are either rare variants or of unknown etiology (14, 28). The apparently high frequency of these abnormalities has been linked to differential mortuary behavior in the Upper (8, 29) and to possibly high levels of consanguinity among Pleistocene humans (14, 30- 32). These assessments of overall patterns have tended to combine abnormalities that are developmental versus degenerative, of clear etiology versus unknown cause, of functional significance versus probably unknown to the individual involved, and/or from associated skeletons versus isolated remains. In this context, and especially with the steadily increasing paleopathological record for Pleistocene humans, it is appropriate to reconsider one general class of abnormalities among these past foraging populations – those that derived from irregularities of developmental processes (other than non- specific stress indicators, which are primarily reflected in the ubiquitous dental enamel hypoplasias). In general, these abnormalities can be divided into those that were systemic versus anatomically localized, to the extent that the distinction can be assessed on incomplete paleontological remains. They can be separated into those for which at least a proximate etiology is evident and those currently of unknown cause, the former being separable into ones whose ultimate etiology is evident versus those which may have multiple causes. The ones that are evidently pathological can be separated from those that appear as unusual and rare variants on “normal” developmental processes, the latter limited to ones that are infrequently found in large recent human samples (and hence would be unexpected in the small available fossils sample sizes for the given element, especially if those samples are limited in time and space) or exceptional in terms of size and/or proportions for the relevant paleontological sample. Given these considerations, following is an annotated review of the currently documented developmental abnormalities among Pleistocene humans, excluding isolated non-specific stress indicators. It is likely that it represents a minimum inventory, given the recent (primarily since the 1990s) documentation and/or reassessment of a substantial number of abnormalities, a number of them noticed only recently on long known specimens. They are grouped in regional anatomical terms, from systemic ones through cranial and mandibular, dental, vertebral, and , pelvis and leg, and hand and foot ones. Within each anatomical group, they are arranged largely alphabetically by sites, independent of the time period/morphological group of the specimen(s) involved, with a few grouped by the form of the anomaly.

Materials and Methods Almost all of the abnormalities presented here have been described in varying levels of detail and many subjected to differential diagnosis. The lesions or variants are therefore summarized, with appropriate references, although the interpretations here do not universally follow those of the original assessments. The descriptions and interpretations are based on the published details, supplemented when possible by personal assessments. Images of the primary abnormalities are provided, as available through personal research and the generosity of colleagues; published figures are referenced for specimens for which images are not available. As appropriate for issues of body, cranial or dental size and proportions, measurements and indices of the remains are compared to relevant samples of Pleistocene human remains. All of the measurements follow the Martin system (33) or Howells (34). Percent asymmetry is calculated as [(maximum – minimum)/minimum] (35). Within the context of anatomical regions, the abnormalities are primarily presented by individual. Eight specimens ( 2, Dolní Věstonice 15 and 16, Kebara 2, Shanidar 3 and 4, Sunghir 2, and Villabruna 1) appear more than once, given more than one apparently independent anomalies. Two additional specimens (Lazaret 18/19 and Pech-de-l’Azé 1) have two abnormalities each, presented together.

2

The specimens are limited to Pleistocene members of the genus Homo, not including initial Pleistocene remains variably included in Homo versus Australopithecus (see 36). In addition, the Middle and sample from and Mata Menge, Flores, (37-39) and the Middle Pleistocene one from Dinaledi and Lesedi, South (40, 41) are not included. The status of each sample relative to other Pleistocene Homo fossils is unresolved, especially since each sample exhibits morphological features that cannot be easily explained phylogenetically in a broader hominid context (42, 43). Many of the abnormalities can be referred to clinically known conditions, even if the precise natures of the defects or are not known. Several of the variants are unique or at least not documented in the clinical or paleopathological literature, and care has been taken to assure that they are abnormalities and not merely morphological variants characteristic of Pleistocene populations. When the anomalies are clinically known and data are available on their incidence in sufficiently large recent human samples, summaries of those data are included. However, it remains unclear whether the incidences in recent human samples are directly applicable to Pleistocene specimens or samples. Many of the variants with known incidences in recent human samples have unknown ultimate etiologies or can result through diverse etiologies; the relevant clinical literature is replete with vague etiologies that invoke genetic causes (without specific variants), inherited predispositions, and/or forms of stress during development. Moreover, the susceptibilities among Pleistocene people may have been different, given contrasting levels of systemic stress, skeletal hypertrophy and/or craniofacial or body proportions. The recent human data are often the only reference available and are therefore employed as available. As a result, the abnormalities for which there are available data, whether they are systemic or individual, are ascribed general levels of probability (<5.0%, <1.0%, <0.1% and <0.01%) based on the incidence data from recent human samples (Table 1). In the few cases in which the unusual aspect is related to size or proportions, the associated probabilities are based on relevant human paleontological comparative distributions, determined from the general paleontological group from which the specimen derives. Ideally, these probabilities would be adjusted for the number of paleontological specimens available and preserving the relevant portions of the skeleton or dentition. However, it is unclear how the reference samples for such adjustments should be delimited (by site, by time period, by geography, by taxonomic group, etc.?). In most cases, the adjustment would not change the assignment to one of the levels of probability, given the modest sample sizes available for most relevant Pleistocene groups.

Systemic Conditions Arene Candide 3 The final Late (, MIS 2) site of Arene Candide in northwestern has yielded ≈20 individuals, two-thirds of which derive from primary burials (32, 44, 45). The Arene Candide 3 partial skeleton retains the cranium and mandible, the L5, the complete pelvis, the right and left humeri, ulnae and fibulae, and one each of the scapulae, radii, femora and tibiae; additional costal, manual and pedal remains are mixed with those of Arene Candide 4 (32). The bones were found intentionally disturbed with portions arranged around the subsequent burial of Arene Candide 2 (32). Based on the pelvic morphology, pubic symphysis metamorphosis, and dental attrition, Arene Candide 3 is a male who died in the late third or fourth decade (16, 30, 32).

3

As described by Formicola (30), the right humerus, the left tibia, both fibulae, and (to a lesser extent) the left ulna and femur of Arene Candide 3 present abnormal or exaggerated curvatures (Fig. S1). In particular, the right humerus (but not the left one) exhibits distinct lateral bowing, whereas the left tibia and especially both fibulae present medial bowing. The left ulna has the normal “S” curve, but with an exaggerated dorsolateral proximal curvature, and the left femur has a pronounced anterior distal curvature. The other long bones exhibit normal levels and patterns of diaphyseal Figure S1. The major long bones of Arene Candide 3, including curvature. The Arene Candide 3 the right humerus in posterior view, the left humerus, right radius remains also exhibit several and both ulnae in anterior view, the left femur and tibiae plus both enthesopathies, in particular a fibulae in anterior view. The scales are separate for the upper and pronounced bony growth on the lingual lower limb bones. Note the marked curvatures of the right mandibular symphysis, small bony humerus, left tibia and the fibulae, the moderately strong spurs on the coracoid process, a curvature of the left ulna, the pronouced lateral midshaft crest with a dorsal hollow on the right humerus, the expansion of the projecting and posteriorly hollow left ulnae coronoid process, and the high femoral neck angle. The deltoid tuberosity, and changes to the distal anterior curvature of the left femur is not evident in anterior left coronoid process and brachialis view. Images courtesy of Vitale S. Sparacello (Université de tuberosity. Bordeaux, Pessac, ) (copyright MIBAC-SABAP Liguria). The long bones of Arene Candide 3 are also noticeably short for a Late Upper Paleolithic male. The humeral maximum length of 271 mm and the femoral bicondylar length of 388.5 mm are at or below the limits of variation of western Eurasian Late Upper Paleolithic males (Fig. S2); they are below all of the values for the other Arene Candide males (humerus: 283 – 315 mm, n = 5; femur: 420 – 465 mm, n = 4). The femoral diameter (45.0 mm) is also modest in size, exceeding only those of Bichon 1 and Villabruna 1. The combination of a relatively short femur and an unexceptional femoral head size produces a head-to- length index (11.6) [reflecting body mass to stature (48, 49)] that is moderately high, with only Arene Candide 2 providing a higher index Figure S2. Appendicular measures of body size for Early/Mid (E/MUP) and Late (LUP) Upper Paleolithic samples, in millimeters. (Fig. S3). The humerofemoral index of AC3: Arene Candide 3; Cus: Cussac L2A; DV15: Dolní Věstonice Arene Candide 3 is unexceptional, but 15; NK2: Nazlet Khater 2. Individual data from: Sládek et al. (46), its crural index of 81.8 at the bottom of Crevecoeur (9), Villotte et al. (47) and Sparacello et al. (32). the Late Upper Paleolithic variation; Comparative data from personal measurement and primary fossil only the earlier Veneri 2 female has a descriptions. lower Upper Paleolithic crural index

4

(Fig. S3), reflecting a rather short Arene Candide 3 tibia in the context of its already short femur.

Figure S3. Proportional comparisons of appendicular remains for Early/Mid (E/MUP) and Late (LUP) Upper Paleolithic samples. For the indices: femur head-length: head diameter / bicondylar length; humerofemoral: humerus maximum versus femoral bicondylar lengths; crural: tibial maximum versus femoral bicondylar lengths. Data sources as in Figure S2. The left humeral and femoral midshafts are unexceptional in terms of the polar moments of area scaled to body mass times length relative to other Upper Paleolithic remains (50, 51). Yet, the femoral neck-shaft angle of 136°, although normal for a recent mechanized population, is substantially above a Late Upper Paleolithic distribution and well above the next highest value (Ohalo 2) (Fig. S3); the high value indicates low levels of reaction forces (and hence activity levels) relative to other Late Upper Paleolithic humans during early development (52, 53). The Arene Candide 3 remains therefore have a combination of irregular diaphyseal curvatures, enthesopathies, small stature but average body mass, a low crural index, and a high neck-shaft angle. The first three aspects have been differentially diagnosed by Formicola (30) as probable reflections of X-linked hypophosphatemic rickets (or hypophosphatemia; XLH), a rare inherited form of vitamin-D resistant rickets (54). It is unclear to what extent the other features reflect secondary effects of his primary abnormalities. Its incidence in recent humans is 1/20,000 or ≈0.005% (55). He was nonetheless active as an adult, if probably less so when immature. Incidence: <0.01%

Cussac L2A Cussac L2A is a largely complete but disarticulated skeleton found within a bear’s nest, within the Mid Upper Paleolithic (, MIS 3) decorated , in southwestern France (47, 56, 57; Fig. S4). Although remaining in situ for conservation reasons, it has been possible to pelvically infer the adult individual’s sex as male (47). The exposed remains do not reveal any abnormalities, but the individual is nonetheless unusually small for a Eurasian Early/Mid Upper Paleolithic male (47). The estimated bicondylar femoral length of ≈447 mm is only Figure S4. The Cussac L2A human remains in situ on the surface within a 1.41 standard deviations below a male mean, but the humeral bear nest. The bones are partially maximum length of ≈300 mm is 2.23 standard deviations disarticulated from post-depositional below a respective mean (Fig. S2). The femoral head movement within the confines of the nest. diameter, predicted from its acetabular height of 50.1 mm Image courtesy of Patrice Courtaud and (≈41.3 mm) falls below the smallest Gravettian male Sébastien Villotte (Université de Bordeaux, (Předmostí 9: 42.0 mm) and 2.29 standard deviations from Pessac, France). the male mean (Fig. S2).

5

The body proportions of Cussac L2A are also at the limits of the known earlier Upper Paleolithic variation. His crural and humerofemoral indices (81.7 and 67.1 respectively; both using femoral bicondylar length), are among the lowest Gravettian values (81.1 for Veneri 2, and 67.4 for Grotte-des- Enfants 5 respectively, both females) (Fig. S3). His femoral head to length index of ≈9.24 is similarly at the low end of the earlier Upper Paleolithic values, exceeded only by that of the female Grotte-des- Enfants 5 (9.14). As documented by Villotte and colleagues (47), the Cussac L2A individual is at or slightly beyond the known limits of Gravettian males in terms of body size and Gravettian males and females in body proportions, and these slightly modified comparative samples reinforce that conclusion. But as noted by them, he was not necessarily pathological. It remains uncertain whether this individual was merely a short Gravettian male or his size and proportions reflect developmental abnormalities; in either case Cussac L2A is unusual for his context. Probability: <1.0%

Dolní Vĕstonice 15 Dolní Vĕstonice 15 is the middle individual from the triple burial (Dolní Vĕstonice 13 to 15) at the earlier Mid Upper Paleolithic (Pavlovian, MIS 3) Dolní Vĕstonice II site in southern Moravia, (46, 58-60). Based on dental maturation and wear, clavicular and vertebral epiphyses, and pubic symphyseal metamorphosis, Dolní Vĕstonice 15 was 20–25 years old at death (61); both his pelvic morphology (62) and aDNA (63) indicate male sex. The majority of the skeletal elements of Dolní Vĕstonice 15 are developmentally normal, if small for a Mid Upper Paleolithic male. However, the right humerus, the left radius and ulna, and the femora exhibit developmental dysplasias (7; Fig. S5). The left humerus of Dolní Vĕstonice 15 is unexceptional, exhibiting only a slight lateral bowing. The stronger right humerus is similarly laterally bowed and has a marked medial angulation of the distal diaphysis, which turned the otherwise normal distal articulation medially. Radiographically there is no evidence of a previous fracture, with even cortical thicknesses through the medial bend. The right humerus is also distinctly longer than the left one (Fig. S5). The humeri are relatively short; the maximum length of the straighter left humerus (301.5 mm) is well below an earlier Upper Paleolithic distribution (Fig. S2). The associated right radius and ulna are normal. The left forearm, however, has evidence of a fracture of the mid-distal ulnar diaphysis with callus formation. The associated radius has a pronounced dorsal angulation of the mid- proximal diaphysis, rather than the usual straight diaphysis in medial view (Fig. S5); its curvature may be secondary to the ulnar fracture. The femora exhibit low neck-shaft angles (111° and 113°), but ones that are nonetheless Figure S5. The abnormally curved long bones of Dolní within the earlier Upper Paleolithic variation Věstonice 15. Left; the right and left humeri in anterior (Fig. S3). Their proximal diaphyses, however, view. Middle: the left ulna in anterior view and the left are abnormally bowed anteriorly, especially on radius in medial view. Right: the right and left femora the right side (Fig. S5). The femora are in lateral view. Not to scale. exceptionally short, substantially below an earlier Upper Paleolithic range (Fig. S2) and

6 even below a Late Upper Paleolithic distribution. The pelvis is similarly petite; its bi-iliac breadth of 234.0 mm is 3.2 standard deviations from a Mid Upper Paleolithic male mean (276.6 ± 13.4 mm, n = 8) (64). Yet, although the femoral head diameter is small, it is within earlier Upper Paleolithic variation; this combination leads to a femoral head-length index (reflecting body mass to stature) of 12.2 (Fig. S3), above those other Upper Paleolithic remains. Despite the modest dimensions of the femora and the pelvis, and the abnormalities of the femora, the more distal lower limbs of Dolní Vĕstonice 15 are normal, from the patellae to the pedal phalanges (65). An abnormal femoral length to normal tibial length comparison results in a crural index (89.7), which lies well above the Gravettian range (Fig. S3). Moreover, assessments of appendicular hypertrophy scaled to his diminutive body size, his degree of humeral asymmetry, and his pattern of upper limb osteoarthritis, indicate an individual who actively participated in subsistence activities (66). Dolní Vĕstonice 15 therefore exhibits small body size combined with right humeral, left forearm, and bi-lateral femoral deformities. Marked dental enamel hypoplasias, especially on the M1s (67), indicate survival of developmental stress. The left forearm deformities are most likely secondary to a distal ulnar fracture, one of the more common fractures, especially from falls (68-71). The humeral and femoral ones, however, are more difficult to diagnose. There have nonetheless been multiple attempts to diagnose his abnormalities as part of a systemic (e.g., 58, 72-74), including hemi-paralysis, rickets, and chondrodysplasia calcificans punctata (CCP). Dolní Vĕstonice 15 does not match the first two conditions, and the last diagnosis (74) requires Dolní Vĕstonice 15 to be female in order to survive to adulthood; the individual is clearly male (62, 63). Moreover, the femora do not exhibit the articular abnormalities associated with CCP. At present, and following Trinkaus and colleagues (7), there does not appear to be a clear diagnosis for Dolní Vĕstonice 15’s lesions, one which would limit the non-traumatic deformities to one humeral and both femoral diaphyses, produce small body size, leave the remainder of the skeleton (especially sub-femoral) unaffected, and permit the individual to have remained physically active. Unknown etiology

Krems-Wachtberg The Mid Upper Paleolithic (Gravettian, MIS 3) site of Krems-Wachtberg, in eastern , yielded the elaborate burial of two newborns, buried side-by-side on their left sides, facing the same direction, with abundant ochre and ivory beads and covered by a mammoth scapula (Fig. 1a in 75). One meter away was the burial of a slightly older infant, 3 months old, also covered in ochre (Fig. 1b in 75). The same dental developmental ages (9–10 months post-conception) and femoral lengths of the two neonates (Krems-Wachtberg 1 and 2) strongly suggest that they were twins, born at full term but either stillborn or deceased very shortly afterwards (75). These remains are among the very few Middle and Upper Paleolithic neonates known (see 5, 32, 76), and Krems-Wachtberg 1-2 represent the only known Pleistocene case of very probable twins. Data on twinning among recent humans provides continental level frequencies between 0.6% and 4.0% and an average frequency ≈2%, although population level frequencies can be modestly outside of this range (77, 78). Based on these frequencies, it is therefore unusual, but not exceptional, for there to have been twins in the Upper Paleolithic. However, these values should be combined with the low probability of having neonatal deaths to both infants and their remains being preserved together. The probability of finding the Krems-Wachtberg “twins” is nonetheless conservatively considered. Incidence: <5.0%

7

Qafzeh 12 The Qafzeh 12 remains, deriving from an early Late Pleistocene (MIS 5) context in northern (5, 79, 80), consist of a partial , facial fragments, a partial dentition, C2 to T1, and the left scapula, humerus, radius and ulna (5). The individual was 3-4 years of age at death, as indicated by dental development, cranial base synchondroses, and vertebral ossification. The dentition, zygomatic bones, and vertebral remains are normal for the age of the child. The neurocranial vault, despite warping and missing portions, is exceptional, as are the long bones of the left arm (5, 81; Fig. S6). The coronal, sagittal and metopic sutures are fully open, which is unexceptional for a 3-4 year old child. Persistent metopic sutures are variably present among recent humans (4.9% ± 4.3%, 34 samples) (82). Among later Pleistocene crania ≥5 years of age with sufficient preserved, none of the other Middle Paleolithic modern humans exhibit persistent metopic sutures (n = 14), but open metopic sutures are present in 11.8% of Early/Mid Upper Paleolithic humans (n = 51) and 20.0% (n = 25) of late (8). However, the Qafzeh 12 bregmatic is completely open, 39 mm wide and 27 mm anteroposterior (5, 81). It has marked digitations on its margins, preserved especially on the left frontal and right parietal bones (Fig. S6). In a pooled series of recent humans, 100% of 3-4 year-olds (n = 14) have complete closure of the bregmatic fontanelle, and all of the 3-7 years (n = 41) have full closure (83). The Qafzeh 12 open fontanelle is associated with a very long and broad frontal bone, which exhibits a prominent eminence especially on the right side, along with an elevated mid- Figure S6. Superior view of the sagittal curvature. Endocranially, the frontal crest and superior Qafzeh 12 anterior cranial vault, sagittal sinus sulcus are displaced left. Posteriorly, the occipital with the frontal squamous portion bone has the superior sagittal sinus sulcus also displaced left, and the anterior and medial parietal producing a large and deep right cerebral fossa. Exocranially these bones. The frontal elongation and changes are reflected in a large and strongly convex superior especially the persistent bregmatic occipital squama. There is also an abundance of lambdoidal sutural fontanelle are evident. Image courtesy of A.-M. Tillier (Université ossicles. The partially preserved cerebellar fossae appear normal. de Bordeaux, Pessac, France). The left humerus, radius and ulna are all markedly gracile, Reprinted with permission from ref. especially as reflected in their thin diaphyseal cortical bone. The 81. humerus is moderately short, especially given the relatively long humeri of the Qafzeh (and Skhul) adults (79, 84), and it exhibits a pronounced lateral bowing, whereas the ulna and radius are rather straight. The combination of cranial features reflects infantile hydrocephaly, in which an excess of cerebrospinal fluid promoted the asymmetrical and expansive neurocranial growth and delayed closure of at least the bregmatic fontanelle (81, 85). The gracility of the left arm may well reflect secondary effects of any neurological deficiencies from the hydrocephaly. In two clinical samples of recent humans, hydrocephaly occurred in only 0.07% (n = 589,510) of live births (86, 87); the probability of finding so clear a case as Qafzeh 12 in the available infantile Late Pleistocene sample is therefore exceedingly low. Incidence: <0.1%

8

Romito 2 The Late Upper Paleolithic (MIS 2) levels in the Grotta del Romito in southern Italy have yielded the remains nine individuals buried either as double burials in the rock shelter in front of the cave (Romito 1 and 2, 5 and 6) or as single burials within the cave (Romito 3, 4, 7–9) (88-90). Although the burials span several millennia (89), they are all associated with Epigravettian assemblages (91). Various of the skeletons exhibit non-specific stress indicators (88), and Romito 8 experienced serious trauma (90). Yet, it is the unusual Romito 2 partial skeleton which has received the most attention (26, 29, 88, 92-95). The largely complete remains of Romito 2, buried alongside the adult female Romito 1, exhibit a suite of skeletal abnormalities related to a form of dwarfism (Fig. S7; see Figs. 6 and 7 in 88 and Figs. 2–5 in 26). Based on a combination of dental (partially formed M3 roots) and varying degrees of synchondrosis and epiphyseal fusion (sphenooccipital synchondrosis almost fused, fused major long bones epiphyses with evidence of the fusion lines, partially fused pelvic epiphyses and unfused clavicular ones), the individual was probably close to the end of the second decade at death (88). The sex of Romito 2 is less certain but was probably male (90). The Romito 2 skeleton is foremost very small. Its humeral maximum length of 168 mm is 55.0% of a Late Upper Paleolithic male mean and 56.4% of a female mean; its femoral bicondylar length of ≈221 mm is 51.1% of a male mean and 53.3% of a female mean; they are 8.5 and 10.9 standard deviations below the male means (see Fig. S2). Relative to the otherwise small Romito 1 female, its humeral length is 64.6% of her length. Its ulnar length (≈111mm) is ≈51.4% of that of Figure S7. The Romito 1 adult female and the Romito 1 (216 mm) and further from the means of Romito 2 adult male in situ, the latter buried on the Late Upper Paleolithic male (259 mm) and female left side of the former. The small size of Romito 2 (248 mm) samples. The tibia length (≈209 mm) is reflects his acromesomelic . Image courtesy modestly less diminutive, being 56.4% and 60.2% of Fabio Martini (copyright: Museo e Istituto of similar male and female means. The tibial and Fiorentino di Preistoria, Florence, Italy). femoral lengths provide a crural index of 94.6, which is substantially above the Upper Paleolithic ranges of variation (Fig. S3). These dimensions are joined by a suite of abnormalities. The cranium has basicranial compression, with a small foramen magnum; the distinctly convex frontal and occipital sagittal contours with prominent frontal and parietal eminences may be related features. As noted, the long bones of the limbs are markedly short, but the partial appear not to be reduced. The epiphyses of the femora and tibiae are largely normal, but they appear disproportionately large for the shortness of the bones. The neck-shaft angle of the left femur (≈136°) is high, similar to that of Arene Candide 3 (see above) and unusual for an Upper Paleolithic human (Fig. S3). The more marked long bone changes are in the cubital regions, the forearms and hands. The proximal ulnae are dysplastic, limiting elbow extension to ≈130°, and lack radial articular facets. The ulnae are strongly bowed dorsally, along with an exaggerated lateral bowing of the left radius. There is also an abbreviation of the metapodials and phalanges with marked dysmorphy of the metacarpals, even

9 though at least the carpals are largely normal. Some of the vertebrae are ventrally abbreviated, suggesting some degree of . This combination of features has been diagnosed as a form of acromesomelic dysplasia (26, 88, 94). There are elements that resemble achondroplasia, but the majority of the deformities and dysplasias fit best with a diagnosis of one of the varieties of acromesomelic dysplasia (96, 97; see discussion in 26). Independent of which specific form of dysplasia affected Romito 2, it is truly exceptional to have an individual with this condition in the limited sample of Late Upper Paleolithic (or Upper Paleolithic more broadly) remains. The incidence of these cases is reported as 1 per 1-2 million live births (26). Incidence: <0.01%

Sunghir 2 The Sunghir 2 skeleton derives from the Early/Mid Upper Paleolithic (MIS 3) site of Sunghir, in northern (98, 99). Sunghir 2 was an early adolescent (≈12 years) male, based on his dentition, pelvis and aDNA (8, 100). He was buried with Sunghir 3 in the richest known Upper Paleolithic burial (8, 99, 101). The Sunghir 2 remains appear superficially normal, and at least the long bone diaphyses had an appropriate level of cortical hypertrophy for a Pleistocene early adolescent (102). However, his dental enamel hypoplasias indicate elevated and especially persistent systemic stress throughout his abbreviated life (103). There are several other aspects that suggest unusual development (8)

Figure S8. Unusual aspects of the Sunghir 2 early adolescent remains. Upper left: lateral view of the facial skeleton showing the weak masticatory muscle entheses, the relatively vertical nasal bridge, and the marked alveolar prognathism. Lower left: the maxillary dentition in occlusal view, showing the almost complete absence of occlusal wear. Middle: the right humerus in anterior view with little to no evidence of the proximal diaphyseal muscle or deltoid insertions. Right: posterior view of the left femur, with its smooth gluteal tuberosity and linea aspera. Not to scale. The normally erupted and occluding dentition (all except the M3s) exhibits essentially no occlusal or interproximal wear, only a narrow band of exposed dentin on the incisor incisal edges and a trivial amount of enamel faceting of the M1s (Figs. S8 and S32). In addition to being exceptional for a Pleistocene forager, this minimal level of occlusal wear contrasts with the high levels of wear evident on the Sunghir 1 and 3 dentitions. The cranium and mandible have little to no marking for the masticatory muscles, only a weak temporal crest that fades out by the ; the masseter and medial pterygoid insertions on the mandibular ramus are smooth. The major postcranial muscle insertions are all weak, with no enlargement of the deltoid tuberosities, slight ridges for pectoralis major, and smooth gluteal tuberosities, linea aspera and soleal lines. Although entheses are only approximate indicators of muscle development, especially in immature individuals (104), those of Sunghir 2 are markedly weaker than those of the younger Sunghir 3 and other Late Pleistocene late juvenile/early adolescent remains. In addition, the facial skeleton provides a profile that is unusual among Upper Paleolithic western Eurasians (Fig. S8). There is little depression of nasion, the nasal bones remain largely vertical, and the subnasal area exhibits marked alveolar prognathism. The last feature superficially resembles the “alveolar prognathism” of some sub-Saharan African populations, but that configuration does not appear there until the terminal Pleistocene, 20,000 years after Sunghir 2.

10

These aspects of Sunghir 2 are not necessarily “pathological” (beyond the pronounced dental enamel hypoplasia), but they provide a morphological pattern otherwise unknown among MIS 3 foragers (facial shape) and levels of dental wear and enthesis development at variance with expected Pleistocene behavioral patterns. It is unclear what developmental abnormalities might have produced these patterns, and whether the unusual of Sunghir 2 (and Sunghir 3 – see below) is related to his elaborate burial (8, 29). Unknown etiology

Craniofacial Conditions Arene Candide 12 – Neurocranium The partial cranial vault of Arene Candide 12 [formerly Arene Candide 19 (32, 45)] presents an unusual elongation of the posterior cranial vault (105). It is part of the Arene Candide 12 skeleton of an adult male, deposited in the stones of grave XV, one of the individuals from the large Late Upper Paleolithic (Epigravettian; MIS 2) human skeletal sample from Arene Candide in northwestern Italy (32, 44, 45). The cranium and mandible of Arene Candide 12 (Fig. S9; see also Fig. 1 in 106) retain the essentially complete facial skeleton and most of the cranial base. The neurocranial vault is present in the supraorbital region, the inferior , and the temporal bones. There is a large left parietooccipital portion, articulating with the temporal bone and including the left . It also includes a left frontoparietal piece and a right inferior parietooccipital section. The preserved portions are sufficient to indicate the overall contours of the neurocranium, if not its full height (Fig. S9). As documented by Formicola and Scarsini (105), the vault would have been relatively low and especially posteriorly elongated. It contrasts with the other Arene Candide crania, which exhibit shorter and more vertically rounded Figure S9. Views of the Arene Candide 12 cranium. A: right lateral; posterior profiles. B: left lateral; C: inferior. Note that Artificial cranial deformation has been suggested for the lateral photographs are taken Arene Candide 12 (106), but it would represent an exceptional from slightly above horizontal. case from the European Upper Paleolithic (105). More relevant, Images courtesy of Vitale S. Sparacello (Université de Bordeaux, however, is that the cranium does not conform to any of the usual Pessac, France) (copyright MIBAC- forms of artificial cranial deformation, and in particular it SABAP Liguria). contrasts with the anteroposterior and circumferential forms (107, 108). It lacks the frontal and occipital flattening of the anteroposterior deformation. It resembles the circumferential form in being posteriorly elongated, but Arene Candide 12 lacks the posterosuperior elevation of the occipital bone with a strongly upward sloping nuchal plane. Its increased length is largely parallel to the Frankfort horizontal. In addition, there is little or no alteration of the facial skeleton, as would be expected from artificial deformation (107, 109). Alternatively, as suggested by Formicola and Scarsini (105), the individual may have experienced premature . The inhibition of growth at the sagittal suture prevents lateral neurocranial expansion and promotes compensatory growth at the coronal and lambdoid sutures (110). All of the sagittal suture is absent. Yet, Arene Candide 12 has the longitudinal and horizontal expansion associated with compensatory coronal and lambdoid sutural growth. Nonetheless, the preserved portions of the frontal bone do not exhibit the bulging associated with compensatory metopic suture growth (110).

11

Therefore, if premature sutural synostosis was responsible for the unusual cranial vault shape of Arene Candide 12, it would have occurred after the cessation of infantile metopic growth but prior to the attainment of near adult brain size by an early juvenile age. The developmental source of the unusual Arene Candide 12 cranial contour is therefore unclear but appears most likely due to restrictions on normal sutural growth from premature synostosis. If that is the case, it would represent a markedly rare condition (0.07% of live births; 0.04% – 0.14% in 7 clinical samples) (111-113), of which about half involve the sagittal suture alone (111, 114). Incidence: <0.1%

Atapuerca-SH cranium 14 – Neurocranium The mid Middle Pleistocene (Acheulian; MIS 11–12 ) site of Sima de los Huesos in the Sierra de Atapuerca (Atapuerca-SH), north-central has yielded thousands of human remains including 17 juvenile to younger adult crania (115, 116). The remains reached the Sima de los Huesos through a long vertical shaft within the karstic system, even though the ultimate processes leading to their descent to the Sima de los Huesos is uncertain (117). Among the human crania, the juvenile Atapuerca-SH cranium 14

Figure S10. Views of the Atapuerca-SH Cranium 14 neurocranium. Upper left: anterior view; lower left: virtual endocast of the posterior endocranial cavity with the mid- cranial slice of the neurocranium; right: superior view. Modified from ref. 25. exhibits pronounced developmental asymmetries (25, 118; Fig. S10). Based on the sphenooccipital and jugular synchondroses (both open), the endocranial capacity (within the adult range), and supraorbital torus size (below those of early adolescents), AT-SH cr. 14 had an age-at- death during the juvenile years (approximately 6–12 years) (25). Sex is indeterminate. The metopic suture is fully fused, and the coronal, sagittal and right lambdoid sutures are open and age appropriate. The left lambdoid suture, however, is almost completely fused (if still evident), all except for ≈31 mm adjacent to lambda. There is also the probable remains of a lambdoid sutural ossicle, one fused to the remainder of the occipital squama inferiorly. In conjunction with these abnormal sutures, the cranium has a pronounced right parietal eminence protruding posteriorly and superiorly, a left occipitomastoid bulge, and an inferior tilt to the posterior cranial base. There is distinct frontal bossing, such as is normally absent from archaic Homo crania. The whole of the cranial base is twisted, including the orientations of the foramen magnum and the temporomandibular articulations relative to the sagittal and/or coronal sutures. These exocranial aspects reflect endocranial proportions (Fig. S10). As differentially diagnosed by Gracia and colleagues (25, 118; see also 110), Atapuerca-SH cr.14 represents a case of premature unilateral lambdoidal synostosis, in which the asymmetries, frontal and parietal eminence protrusions, and cranial base twisting would be secondary growth compensations for the inability of the left lambdoid suture to accommodate the apparently normal brain volumetric growth [its endocranial capacity of ≈1,224 cc is well within the Atapuerca-SH adolescent-to-adult range (115)]. In one large modern human sample, unilateral lambdoidal premature synostosis had a prevalence of 0.001% of live births (113), making the probability of finding such an individual (even with the two dozen sufficiently complete European mid Middle Pleistocene crania known) vanishingly small. Incidence: <0.01%

12

Dolní Vĕstonice 16 – Maxillae The site of Dolní Věstonice II, in southern Moravia, Czech Republic, has yielded a triple burial (see Dolní Věstonice 15 above) and the isolated burial of Dolní Věstonice 16 (60, 119). Cranially, pelvically and aDNA sexed and aged based on dental wear, auricular surface metamorphosis and pervasive vertebral osteoarthritis, Dolní Věstonice 16 represents an older (fifth decade) male (61-63). The burial is associated with an earlier Mid Upper Paleolithic (Pavlovian) assemblage (MIS 3). The cranial abnormality of Dolní Věstonice 16 consists of a misalignment of the maxillae along the intermaxillary suture (120; Fig. S11). The palate is present laterally close to the maxillopalatine sutures, but the intermaxillary suture is present only from just posterior of the incisive foramen to the labial alveoli. The right side of the palate is inferiorly positioned ≈4 mm relative to the left one anteriorly, and the left palate is modestly inferior of the right one posteriorly. In addition, the left maxilla is ≈4 mm anterior of the right one, as is evident in the I1-I1 interdental alveolar bone (Fig. S11). There is also remodeling of the anterior nasal spine, such that it is obliquely oriented. Given the damage to and reconstruction of the Dolní Věstonice 16 cranium and the addition of wax filler, one could query whether the misalignment of the maxillae relative to each other is a post-mortem/restoration . However, the bones are naturally (ante-mortem) fused to each other between the I1 alveoli. More importantly, the heavily worn Figure S11. The Dolní Věstonice 16 maxillae in anterior (above) and dentition shows an even wear occlusal (below) views. Note the superoinferior and anteroposterior and a normal occlusion with displacements of the right and left maxillae along the intermaxillary suture, the complete and undistorted and the ante-mortem distortion of the anterior nasal spine. Occlusal image courtesy of John C. Willman (Institut Català de Paleoecologia Humana i mandible (67, 121). The Evolució Social, Tarragona, Spain). misalignment was therefore present long before the death of Dolní Věstonice 16, and most likely long before the completed formation/eruption of his permanent dentition. Vlček (72) suggested that the Dolní Věstonice 16 abnormality was the result of a facial fracture, which separated the two maxillae along the intermaxillary suture. In order for this to have occurred, it would have necessitated forcibly separating the two maxillae, without serious damage to the adjacent portions of the facial skeleton. However, the intermaxillary and especially interpremaxillary sutures are formed prenatally, and growth at these sutures ceases 1 to 2 years postnatal (122). The facial skeleton of Dolní Věstonice 16 was damaged and has been restored to some extent. There are also minor/trivial traumatic lesions on the cranial vault (7), as there are commonly on Pleistocene (including Dolní Věstonice) human remains (7, 13, 72). Yet, evidence of trauma or post-traumatic remodeling, from an insult sufficient to separate and realign the maxillae, should be evident in the preserved portions, and it is absent. The alternative interpretation is a developmental malformation of the intermaxillary area (cleft palate), with a resultant misalignment of the maxillae prenatally, during the formation of the interpremaxillary and intermaxillary sutures (7, 120). Damage to the palatal midline posterior of the incisive foramen area precludes knowing whether the more posterior palatal midline was patent, as often occurs in cleft palate. It is therefore likely that Dolní Věstonice 16 sustained at least a minor form of

13 developmental intermaxillary misalignment. Cleft palate is relatively rare among modern humans, occurring in an overall incidence of 0.16% (0.02% - 0.40%; 74 worldwide samples) (123). Incidence: <1.0%

Palomas 6 and 23 – Mandibles The late Middle Paleolithic (MIS 3) site of Sima de las Palomas del Cabezo Gordo, in southeastern Spain, has yielded >100 Neandertal elements (mostly teeth, but including three partial skeletons) (124-126). A substantial portion of the human sample was found ex situ on the adjacent slope, discarded by 19th century miners, and among those remains were two partial mandibular corpori, Palomas 6 and 23 (Figs. S12 and S13). Although out of context, the Neandertal morphology and probable original stratigraphic locations of the mandibles indicate a Middle Paleolithic, MIS 3 age (127, 128). Both of these mandibles exhibit an unusual addition of bone along the anterior and lateral margins of the inferior corpus.

Figure S12. Anterolateral (A), inferior (B) and occlusal (C) views of the Palomas 23 right mandibular corpus. The white arrows indicate the extents of the extra inferoanterior and inferolateral extra bone, which has become partially separated from the normal corpus anteriorly from heating. The medial seam of the new bone below the molars is also indicated by the yellow arrows. The mandibular corpus of Palomas 23, from the mid- symphysis to the right gonion, is unilaterally complete. Extending anteriorly from below the distal M3 to the area of the M1, there is an extra strip of bone largely fused to the inferior lateral corpus. From the M1 around to the mid-symphysis, the extra bone continues on the lateral edge of the inferior margin, lying outside the normal right digastric fossa. It projects laterally in the vicinity of the P4-M1, forming a slight angle in lateral view (Fig. S12A and S12C). Along its anterolateral and anterior portions, it became separated from the corpus, apparently from post-mortem heating in situ, but its line of adherence to the corpus more posteriorly is readily apparent in inferior view (Fig. S12B). The line of attachment, or partial fusion, of the additional bone to the corpus is also evident in μCT sections of the corpus (128). The Palomas 6 left mandibular corpus is intact from the symphysis to the middle of the M3, although the inferolateral margin sustained minor erosion (Fig. S13). It presents an everted ridge of bone along the anterior and lateral inferior margins, from the symphyseal break to the M2. It is most prominent below the C1 to P4, tapering off posteriorly and medially Figure S13. Anterolateral (A), across the symphysis. The ridge was inferior (B) and anterior (C) probably distinctly more projecting views of the Palomas 6 left prior to the post-mortem erosion. It mandibular corpus. The arrows is fully fused to the mandibular indicate the extent of the corpus and not separable as with the inferolateral and inferoanterior additional bone on Palomas 23. “flange” of bone. It is unclear what these “flanges” on the inferior mandibular corpus bone represent. Similar mandibular morphology is unknown among both Neandertals and modern humans. This uniqueness applies especially to the strip of bone on Palomas 23 with its separate ossification and then partial fusion

14 to the normal corpus. The associated corpus margin is usually rounded from lateral to inferior with a variably pronounced anterior marginal tubercle (129). Although an association between these “flanges” and anterior marginal tubercles has been considered (128), they bear little resemblance to that feature. Anterior marginal tubercles occur from P3/P4 to M1 among the Neandertals (129); the more prominent portion on Palomas 6 is in that region, but the one Palomas 23 is more mesial. Moreover, an anterior marginal tubercle is a discrete protuberance and not a long everted area of bone from symphysis to the distal molars. And there is no separation of anterior marginal tubercles from the corpus surface, as there is with Palomas 23 (if not Palomas 6). These aspects of these two mandibles therefore represent an unusual, and apparently unique, feature. It is not present on the Palomas 1 and 59 mature mandibles, although there is a suggestion of an incipient one on the Palomas 7 ≈4 year-old mandible. It is unlikely to have had any functional consequence for the individuals. Unknown etiologies

Pech-de-l’Azé 1 – Cranium The child’s from the site of Pech de l’Azé I in southwestern France for a long time provided one of the few reasonably complete facial skeletons of a very young Neandertal child (Fig. S14). Excavated in 1909 (130), first described half a century later (131), and securely dated to the late Middle Paleolithic (MIS 3) an additional half century later (132), it is now joined by a number of young Neandertal children retaining variable portions of their facial skeletons (e.g., 133-138). It nonetheless has remained central to considerations of Neandertal craniofacial ontogeny and its variability (e.g., 134, 139, 140). These considerations have assumed that the Pech de l’Azé 1 craniofacial remains represent a normal pattern of Neandertal early development. However, there are two developmental aspects of its cranium that are at or beyond the limits of at least recent human variation, features unrelated to its Neandertal affinities. The Pech de l’Azé 1 skull retains a largely complete deciduous dentition with the developing crowns of most of the more mesial (distally to the M2s) permanent teeth (Fig. S14D). It has been aged to ≈26 months by Patte (131) and to ≤2 years by Legoux (141, 142), although both considered ages-at-death up to 2.5 years (see also 143). Comparisons of its degree of dental calcification to the modal patterns of Ubelaker (144) and AlQahtani et al. (145) provide an age of 2.5 years or greater (see also 146). The degrees of fusion of its temporal and mandibular synchondroses provide ages within this range

Figure S14. The skull of Pech de l’Azé 1 in superior (A), right lateral (B), and left lateral (C) views, with an occlusal view of the maxillae (D), not to scale. The arrows in the lateral views indicate the inferior extents of the mastoid processes. (131, 143, 146). Its fully patent metopic suture is not relevant to age assessment, given its frequent persistence in recent humans (82) and on 12.2% (n = 90) of Middle and earlier Upper Paleolithic crania ≥5 years of age (see above).

15

The anterior cranial vault of Pech de l’Azé 1 exhibits a large bregmatic lacuna (Fig. S14A), which is partially due to post-mortem loss of bone, especially on the left side. However, along the posteromedial margin of the right frontal bone, there is a thinned margin which represents the border of a large anterior (bregmatic) fontanelle. The preserved edge of the fontanelle is 13 to 15 mm long (131, 143), indicating that the fontanelle was only moderately reduced from its maximum dimensions. Anterior are normally completely closed by the middle of the second year (147), may remain partially open into the third year, but are almost always fully closed by the end of that year (83). Given a modal age-at-death of ≈2.5 years for Pech de l’Azé 1, it would be unusual but not exceptional for it to retain some evidence of the anterior fontanelle. However, the persistence of a large open fontanelle into the middle of the third year postnatal would be very unusual, especially because the data on closure by age (e.g., 83) indicate the complete ossification of the area of bregma. The persistence of a large anterior fontanelle in Pech de l’Azé 1, therefore, can only be considered normal if its age-at-death is minimized and the potential age of large fontanelle presence is maximized. It should be noted that these considerations hold even if the rate of development of Pech de l’Azé 1 (or other Neandertals), as based on their dentitions, was faster than the recent human reference samples employed for dental calcification aging (see discussion in 148). If dental development is a good reflection of Neandertal overall developmental rates, then cranial vault ossification should have been accelerated to the same degree as dental formation might have been. Pech de l’Azé 1 would still exhibit delayed anterior fontanelle closure. The other unusual aspect of Pech de l’Azé 1 concerns the asymmetry of its mastoid processes. Both of them are only slightly projecting laterally and curve inferomedially, as is age appropriate. The right one extends to the level of the inferior auditory porus (Fig. S14B), whereas the left one projects below the level of the lateral tympanic bone (Fig. S14C). The resultant mastoid height measurements, from the Frankfort plane (Howells (34) MDH), are 10.5 and 14.5 mm, which provide an asymmetry value of 38.1%. This value is seven times the median (4.7%, n = 50) of a geographically mixed sample of recent human adult crania with a highly skewed distribution of asymmetry values, and it is 65.6% higher than the maximum value (23.0%) in that sample. It is also far from the value of 5.6% for the ≈3 year-old Engis 2 Neandertal and from those of 1.5% for a 2-3 year recent human and 2.2% and 4.3% for two recent human 10–12 year olds; this immature sample is small, but it should be sufficient to indicate that mastoid height asymmetry should be low through development. The mastoid height asymmetry of Pech de l’Azé 1 is therefore exceptional, in a portion of the cranium (the cranial base) that tends to be developmentally conservative. Therefore, although the Pech de l’Azé 1 skull appears to be generally representative of Neandertal third year developmental morphology, given normal inter-individual and age-related variation (e.g., 134, 140, 143), its late persistence of a large bregmatic fontanelle suggests ontogenetic delays which may have affected other aspects of its development. The asymmetry of its mastoid processes may be merely a feature exceeding normal human variation, or it may reflect other unusual aspects of its cranial base. Probabilities: <1.0% and <0.01%

Rochereil 3 – Neurocranium The Grotte de in southwestern France yielded Late Upper Paleolithic ( and , MIS 2) levels, with human remains deriving from both horizons (149-151). From the Magdalenian level and directly dated to the terminal Pleistocene is the largely complete if fragmentary skull of a child, Rochereil 3 (152). Rochereil 3 was originally poorly reconstructed (Fig. S15A) and described (150) as hydrocephalic and having sustained a post-mortem frontal trephination. Re- reconstructed and reassessed (152, 153; Figs. S15C and S15D), the immature Rochereil 3 remains present

16 a set of unusual features. Its dentition (Fig. S15B) has a fully erupted deciduous dentition with little wear, two-thirds of the incisor crowns, half of the maxillary canine crown, and the complete M1 crowns, suggesting an age of ≈3 years or 2–4 years (152; see also 144, 145). The original diagnosis of hydrocephaly is currently untenable, based on the normal formation of the cranial vault sutures and the complete closure of the anterior (bregmatic) fontanelle. Mafart et al. (152) retained a descriptive diagnosis of “macrocranium” but only if the child was female and 2 years old. Their estimated endocranial capacity of 1417 ± 92 cc is indeed large for recent human child of ±3 years. However, if its mean volume estimate is converted to brain weight following Martin (154) (1370.2 gm), then converted to a percentage of adult brain weight for European Late Upper Paleolithic humans (1417.5 ± 107.2 gm, n = 22), and then the latter value compared to a 20th century human sample (155; see also 156), the endocranial volume of Rochereil 3 becomes only moderately large. For 2, 3 and 4 year olds respectively, the Rochereil 3 percent of mean adult brain weight (96.6%) is 2.13, 1.25 and 0.77 standard deviations from the recent human age- specific means. In other words, as noted (152), as a Figure S15. The Rochereil 3 immature two-year-old, Rochereil 3 would have exhibited some craniofacial remains. A: the cranium as preserved; degree of macrocrania, but as a more likely three- B: occlusal view of the maxilla; C and D: virtual year-old child, it would have been unexceptional. reconstruction of the cranium (in gold), with the right and symphyseal portions of the mandible (in At the same time, the frontal bone and blue) and a mirror image of the corpus for the left mandibular corpus exhibit curious osseous lacunae. side (in yellow). Not to scale. Note that the The large circular void in the frontal squamous mandibular anterior teeth are not human (153). CT portion (40-45 mm in diameter) preserves the inferior images courtesy of Gaspard Guipert (Collège and right margins intact, which are smoothed with Lycée Pastré-Grande Bastide, Marseille, France). microporosity, thin to an acute angle, lack exposed diploë, and have marginal sclerosis of the bone (152). The surrounding internal and external tables are normal. It is therefore a remarkably round and large cranial lytic lacuna. In addition, the right mandibular corpus exhibits an oval lacuna with dense margins below the deciduous canine, which is connected with a narrowed inferior alveolar nerve canal. Mafart and colleagues (152) considered a suite of conditions to account for the cranial lacuna (with or without the mandibular one), including fibrous dysplasia, epidermoid cysts, neurofibromatosis, and Langerhans histiocytosis. None of them match the suite of normal and abnormal bone present on the Rochereil 3 skull, and all of them are rare. A conclusive diagnosis for the abnormalities of Rochereil 3 therefore does not exist. Unknown etiology

Salé 1 – Nuchal Region The mid Middle Pleistocene (MIS 11 or 13) site of Salé in northwestern yielded the neurocranium and maxillary dentition of a young adult from a fossil bearing dune (157). The neurocranium (Fig. S16) retains most of the posterior and superior vault, a more complete left side, and much of the cranial base. The specimen presents a suite of unusual features, especially when compared to other mid Middle Pleistocene crania (158).

17

Figure S16. Posterior (above) The occipital region of Salé 1 and right lateral (below) views lacks any evidence of the strong nuchal of the Salé 1 cranium. Note the torus ubiquitous among irregular nuchal plane and the contemporaneous crania, with an absence of a clear nuchal absence of an external occipital torus. Photographs taken by protuberance, little indication of the the late Chester Tarka insertions for the semispinalis capitis (Ancestors Project, American and trapezius muscles, and an Museum of Natural History, asymmetric undulation of the region of New York, NY); images courtesy of Jean-Jacques the superior nuchal line. The inferior Hublin (Max Planck Institute nuchal line is irregular and for Evolutionary asymmetrical, with the muscle Anthropology, Leipzig, insertions being more pronounced on ), and Eric Delson the right side. In contrast, the right (American Museum of Natural temporal line is gracile on the parietal History, New York, NY). bone relative to the left one. The cranial base is asymmetrical, with the trans- auriculare line being further anterior on the left side, using the sagittal suture to indicate the mid-. The mastoid process is more robust and projects further inferiorly on the left side, and the anterior cranial base has more developed muscular insertions (especially for the lateral pterygoid muscle) on the right side. In addition, the occipital condyles are flattened, and project little from the inferior plane of the occipital bone. This combination of features has been appropriately diagnosed by Hublin (158; see also 159, 160) as representing infantile muscular (CMT), a condition which usually develops from a shortened sternocleidomastoideus muscle but can affect a variety of craniocervical muscles and produce asymmetries of the neurocranium, as well as of the facial skeleton (161, 162). It can have various proximate etiologies, and although CMT is often the result of difficulties with parturition, especially in cases of breach births, it can develop in utero (162-164). Mild forms of torticollis can often be corrected, but it is evident that Salé 1 experienced a marked form of the deformity with little or no amelioration during its three decades of life. As noted by Hublin (158, 159), it is likely to have seriously affected the individual’s function. The incidence of congenital muscular torticollis in extant human populations has been reported as ranging from 0.008% to 2.1% for seven clinical samples (average: 0.9%) (163, 164), and an adult cranial sample provided a frequency of 1.8% (165). Although the values across the studies are not directly comparable, given variations in etiologies, the severity of the CMT, whether the deformities were present at birth, and variation in diagnositc criteria (161), the overall average incidences should provide an appropriate reference for Salé 1 given its young adult age and the pronounced cranial effects of the developmental abnormality. Incidence: <1.0%

Singa 1 – Neurocranium The adult Singa 1 neurocranium from southern Sudan, dated to the late Middle Pleistocene (MIS 6) (166), has presented a series of “archaic” and “modern” features that have made its morphological affinities ambiguous (e.g., 22, 167-172). It exhibits some archaic features in the supraorbital and temporal regions, but the overall shape of the neurocranium contrasts with most non-modern Middle and Late Pleistocene crania in being broader and sagittally rounder, especially posteriorly (Fig. S17). It is also

18 divergent from Later and Holocene African crania. However, as suggested by Brothwell (169) and further assessed by Spoor et al. (22), the unusual shape of the neurocranium is clearly pathological. Exocranially, the Singa 1 neurocranium is characterized by mid-sagittal frontal, parietal and occipital curvatures that are unexceptional for a later Pleistocene human, based on separate arc-chord indices. The frontal and occipital mid-sagittal chords (119 and 104 mm) are well within normal ranges for a later Pleistocene human. However, the bregma-lambda length of the parietal bones (96 mm), as noted by Brothwell (169), is short absolutely and relative to either the frontal (nasion-bregma) or the occipital (lambda-opisthion) lengths (Fig. S18). An index of parietal chord to the frontal one is 80.7 for Singa 1, 2.92 and 3.20 standard deviations from Late Pleistocene archaic and modern human samples respectively and 6.1 standard deviations an African later Middle Pleistocene sample. Similarly, an index of the parietal chord to the occipital one for Singa 1 (92.3) is 3.05 and 2.85 standard deviations from the same samples and far from the value of 135.3 for Irhoud 2. The closest later Pleistocene values are 88.2 (Feldhofer 1) and 104.3 (Mladeč 1) respectively for the two indices. The Singa 1 cranial vault sagittal bone proportions are therefore highly divergent from those of other later Middle and Late Pleistocene humans. These proportions occur with the context of an endocranial volume of 1340 cc (171), which is modest but well within the range of a sample of late Middle and Late Pleistocene archaic humans Figure S17. Superior (above) and (1443 ± 177 cc, n = 21) and above that of Laetoli 18 (1237 cc). right lateral (below) views of the The Singa 1 cranium, in superior view (Fig. S17), Singa 1 neurocranium. becomes markedly broader from the anterior frontal squamous to the mid-parietal bones, in a straight line rather than the gently laterally convex curvature of other Pleistocene crania. It then angles medially at the parietal eminences to its evenly rounded occipital profile. Its maximum cranial breadth of 155 mm is at the top of the Late Pleistocene variation, matched only by the large Amud 1, Herto 1, 1 and Mladeč 6 crania (all ≤157 mm), despite its modest endocranial capacity. That maximum dimension is at the parietal eminences. The unusual shape of the Singa 1 neurocranium, with its relatively short sagittal suture and protruding parietal profile (Fig. S17), appears most likely to have been the product of premature sutural synostosis (110). Given the general exocranial obliteration of the sutures on Singa 1, it is not possible to assess directly any differential fusion of the metopic, coronal, sagittal or lambdoid sutures. Yet. the largely normal frontal bone lacking a median bulge argues against early metopic sutural closing, as does the absence of posterior cranial elongation mitigate against premature sagittal suture fusion (173; as noted by 25). The largely symmetrical lateral parietal bulging agrees best with bilateral premature coronal suture synostosis, which results in lateral compensatory cranial vault growth and reduced anteroposterior parietal (hence sagittal suture) elongation. Singa 1 lacks the mid-frontal bulging often associated with infantile bilateral coronal suture fusion (110), suggesting that it occurred after the metopic suture was able to respond to from brain growth. Cranial suture synostoses are well-known clinically, either as isolated abnormalities or as components of systemic conditions. Yet, they are rare in recent human samples, occurring in an average of 0.07% (0.04% – 0.14% in 7 clinical samples) (111-113). Of these synostoses, a minority of them involve the coronal suture (111, 113, 114). The probability of finding a cranium with sutural synostosis,

19 and especially bi-coronal synostosis (assuming that this is the pattern that affected Singa 1), among later Middle Pleistocene humans is exceeding low.

Figure S18. Size and proportions of the Singa 1 parietal (bregma-lambda) chord versus later Pleistocene crania. Parietal/frontal index = (parietal chord / frontal (nasion-bregma) chord) x 100; Parietal/occipital index: (parietal chord / occipital (lambda-opisthion) chord) x 100. Africa: later Middle Pleistocene humans from Herto, Irhoud, Laetoli and Omo-Kibish; Nean: western Eurasian late archaic humans; EMH: western Eurasian Late Pleistocene Middle (purple) and earlier At the same time, the parietal eminence prominence of Singa Upper Paleolithic (green) 1 was exaggerated by a bilateral diploic expansion of the supero- modern humans. The Singa 1 are middles of the parietal bones (22, 171). Although Singa 1 does not data from Brothwell (169), and show evidence of the diploë having expanded through the external the comparative data are from table, the major expansion of the parietal (but not frontal) diploë is primary descriptions and commensurate with an expansion of the erythropoietic marrow within personal measurement. the parietal bones (22). Often considered a reflection of iron deficiency anemia, this condition is better associated with hemolytic or megaloblastic anemias (174). It may also be related, in Singa 1, with the circulatory abnormalities inferred to have been associated with its unilateral (right) temporal labyrinth ossification (22). The unusual overall shape of the Singa 1 neurocranium therefore appears to be a consequence of several factors. The individual had a premature sutural synostosis, apparently of the bi-coronal suture, leading to a diminutive mid-sagittal arc and lateral parietal expansion. A further exaggeration of the parietal eminences was due to diploic expansion of the kind associated with moderate erythropoietic marrow hypertrophy. The unilateral labyrinthine ossification, of unclear but probably circulatory etiology, would have engendered partial deafness, vertigo, postural problems and probably other neurological deficiencies (22); it may have been etiologically associated with the diploic expansion. The Singa 1 individual therefore sustained developmental abnormalities related to the sutural synostosis and further complications related to the labyrinthine and diploic alterations. Incidence: <0.1%

Xujiayao 11 – Parietal Bones The site of Xujiayao, in the Nihewan Basin of northern , yielded an abundant large mammalian , a late lithic assemblage, and 17 archaic human cranial, mandibular and dental remains (175-178). The archeological-paleontological levels are early Late Pleistocene in age (e.g., 179-180); recent later Middle Pleistocene (MIS 7–9) dates (181) are not securely associated with the paleoanthropological remains. The cranial and dental remains have their closest affinities with earlier and roughly contemporaneous eastern Asian Middle and Late Pleistocene remains (180, 182, 183), although Xujiayao 14 and 15 exhibit a mix of archaic, “Neandertal” and more recent attributes (178, 184). Among the cranial remains, however, a bilateral set of parietal bones (Xujiayao 11) exhibits a rare variant, an enlarged parietal foramen (14; Fig. S19).

20

Xujiayao 11 retains 41 mm of the posterior sagittal suture with a large Pacchonian depression on more anterior portion of the right side, as well as two smaller depressions more posteriorly. There is no evidence of meningeal sulci, or the lambdoid suture posteriorly. The pieces therefore represent the middle to posterior portions of the right and left parietal bones, along a fused and largely obliterated sagittal suture. The tables are thin with expanded fine diploë. It is therefore from an older adult. It is notable for a lacuna on the posteromedial portion of the right . The left portion of the opening is the result of post-mortem breakage and reassembly of the pieces of vault bone, but the right portions were rounded ante- mortem (Fig. S19). The opening is 9.4–10.4 mm sagittally and estimated at 10.8–13.3 mm coronally (14). Alongside of the posterior right-deviated sagittal suture, endocranially there is a moderately deep sulcus, 12 mm in length and 7–8 mm wide, which extends from the posterior edge of the lacuna to the posterior right margin of the parietal piece (Fig. S19). The sulcus is deepest at the edge of the lacuna, and it then approaches the endocranial surface at the broken edge of the bone. The sulcus could be related to the sagittal sinus, but it is Figure S19. The posterior parietal bones of Xujiayao 11 in exocranial (left) and endocranial more likely to have been for a vessel through the lacuna (right) views, with the area of the enlarged and to a vascular structure within the neurocranium. parietal foramen enlarged below. The black Differential diagnosis of the lacuna (14) makes arrows delimit the margins of the rounded lacuna it possible to exclude trauma, neoplasms, infectious on the right parietal bone. The blue arrow points and porotic hyperostosis as the cause of the to the posterolateral endocranial vascular sulcus. Images courtesy of Xiujie Wu (Institute of lacuna with its vascular sulcus. The most likely etiology Vertebrate Paleontology and Paleoanthropology, is an enlarge parietal foramen (foramina parietalia Beijing, China). permagma), a rare disorder involving abnormal bone development of the skull often resulting in other abnormalities (185). They derive from a malformation of the parietal bones, in which normal symmetrical fetal openings in the parietal bones fail to close (186). They assume various forms and are in the vicinity of the normal parietal foramina (187, 188). The edges are often smoothly rounded at the expense of the outer table. They may be asymptomatic, but they are often associated with cerebral venous anomalies, irregular suture fusion and deviations of the sagittal suture. Enlarged parietal foramina have a recent human incidence of ≈1 per 25,000 (185). They are absent from the Xujiayao 5 and 6b parietal bones and from the eight other sufficiently preserved Middle and initial Late Pleistocene crania from eastern Eurasia (18 if the Ngandong and Narmada crania are included). Nor are they present among the 29 sufficiently intact western Eurasian and African Middle Pleistocene crania or the 22 western Late Pleistocene archaic humans with parietal bones. It is therefore exceptional to find evidence of one in the available eastern Eurasian (or pan-Old World) Middle Pleistocene sample. Incidence: <0.01%

21

Dental Abnormalities Dolní Vĕstonice 15 – Premolars The Dolní Vĕstonice 15 largely complete, young adult male skeleton, the middle one in the earlier Mid Upper Paleolithic (Pavlovian; MIS 3) triple burial from Dolní Vĕstonice II site (southern Moravia, Czech Republic), sustained a series of developmental dysplasias, minor trauma, dental enamel hypoplasias and orthodontic issues (7, 17, 60, 67; see above). In addition to these developmental insults, radiographic survey of his dentition (67) revealed a mandibular supernumerary tooth (Fig. S20). The Dolní Vĕstonice 15 dentition retains 30 of its permanent teeth, the left mandibular incisors having been lost post-mortem (Figs. S20C and S20D). It has crowding of the maxillary dentition, with labial displacements of the canines, twisting of the I1s, and incomplete eruption of the M3s. The mandibular dentition has rotation of the left P4, partial impaction of the left M3, and marked impaction of the right M3. Apparently unrelated to these orthodontic issues, none of which are exceptional in an Upper Paleolithic context (17, 67), is a supernumerary tooth within the left mandibular corpus (Fig. S20A). Evident radiographically, the extra tooth appears as a simple conical crown adjacent to the left C1 and P3 roots. The crown is completely calcified, and there is the initiation of the root. It is unclear whether it would have Figure S20. The supernumerary mandibular continued to form, and achieve eruption, had Dolní tooth of Dolní Věstonice 15. A: buccolingual Vĕstonice 15 survived longer; all of the other teeth radiograph of the mandibular left C1, P3 and (including the M3s) were completely formed with their supernumerary conical tooth (blue arrow); image root apices closed (61). Its presence does not seem to courtesy of Sarah A. Lacy (California State University-Dominguez Hills, Carson, CA). B: have affected the left C1 or P3, both of which erupted normally (Fig. S20D). buccal view of the more mesial left mandibular dentition, from C1 to M1, showing the normal Supernumerary teeth in the permanent eruption of the teeth and the dental enamel dentitions are relatively common in recent human hypoplasias of the C1 and M1. C and D: occlusal populations, occurring in 0.1% - 3.8% of individuals views of the maxillary and mandibular (189-193). The incidence of them in mandibular dentitions; images courtesy of John C. Willman (Institut Català de Paleoecologia Humana i premolars is often second after maxillary anterior teeth, Evolució Social, Tarragona, Spain). The but the resultant frequencies of them in the samples maxillary dental crowding and the incomplete remain generally <0.5% (191-193; but see 194, 195). eruption of the M3s are evident. Not to scale. Non-descript conical ones are also the most common form (192, 193). In addition, supernumerary teeth are consistently more common in males (190, 192). The presence of a conical supernumerary tooth in the mandibular mesial premolar area of the young adult male Dolní Vĕstonice 15 is therefore unusual but not exceptional. Incidence: <1.0%

Dolní Vĕstonice 33 and Pavlov 21 – Anomalous Teeth In addition to the clear supernumerary tooth of Dolní Vĕstonice 15, an additional isolated tooth from the Dolní Vĕstonice II site (Dolní Vĕstonice 33) and one from the adjacent and contemporaneous Pavlov I site (Pavlov 21) (both in southern Moravia, Czech Republic) most likely represent

22 supernumerary teeth (67; Figs. S21 and S22). Each one is small, unusual in morphology, but isolated and therefore out of alveolar context. Dolní Vĕstonice 33 (Fig. 11.11 in 67; Fig. S21) is a small (mesiodistal: 8.1 mm; buccolingual: 7.6 mm) tooth with one larger cusp and two smaller and blended cusps separated from the larger one by a central fissure. The root has multiple portions, fused together. It was unerupted, as indicated by the absence of any crown abrasion. It could represent a very reduced M3, most likely a maxillary one. Figure S21. Occlusal (left) Alternatively, it could be a supernumerary tooth, in which case it and side (right) views of the 4 Dolní Věstonice 33 anomalous would be a distomolar (or M ) given its molariform shape. Fourth molar, probably a molars are relatively rare, occurring in 0.2% and 0.5% in two recent supernumerary distomolar. human samples (191, 192). Images courtesy of Sandra Pavlov 21 (Fig. 11.24 in 67) is small tooth (6.8 x 6.5 mm) with Sázelová (Czech Academy of Sciences, Brno and Mazaryk a globular crown (Fig. S22). It was fully erupted, as is indicated by an University, Brno, Czech occlusal wear facet with a spot of dentin exposure and Republic). hypercementosis of the distal root. The apex of the root also turns at a right angle to the rest of the root. Its crown morphology is most similar to that of a mandibular mesial premolar. However, it has a double interproximal wear facet, with a central ridge separating it into halves.

The interproximal facet indicates that it was not in normal occlusion, but Figure S22. Occlusal (left) was pressed against two teeth at their juncture, with Pavlov 21 against the and interproximal (right) mesial and distal of those teeth, either lingually or buccally. It views of the Pavlov 21 could represent a mandibular premolar which erupted adjacent to the anomalous tooth, either a neighboring teeth (C1 and P4, or P3 and M1). Alternatively, it conforms misaligned mandibular better to being a supernumerary tooth, in which case it would be a premolar or (more likely) a paramolar. As such it would have been buccal or lingual to the paramolar. The yellow arrow interproximal space of adjacent molars, with its root in the adjacent indicates the “interproximal (toothpick) groove”; the alveolar bone. Paramolars (along with canines) are the rarest of double interproximal facets supernumerary teeth and are more commonly mandibular (196, 197). are adjacent on the crown. They are usually noted as isolated clinical cases, given their scarcity, and Images courtesy of Sandra in at least one large sample they occurred in 0.06% (192). Sázelová (Czech Academy of Pavlov 21 also has a transverse groove on the root adjacent to the Sciences, Brno and Mazaryk University, Brno, Czech cervix, just below the interproximal facets. The groove has scratches Republic). running perpendicular to the root axis, and therefore represents an “interproximal (toothpick) groove” (198), apparently from efforts by the individual to remove debris from between the tooth and the adjacent ones. Dolní Vĕstonice 33 and Pavlov 21 would therefore be morphologically very unusual as members of regular permanent dentitions, as would be the interproximal wear pattern on Pavlov 21. Although it is difficult to confirm for isolated teeth, they are best viewed as supernumerary teeth, a distomolar for the former and a paramolar for the latter. Incidences: <1.0% and <0.1%

Garba IV-E43 – Dentition Level E of the Early Pleistocene locality of Garba IV, in the Melka Kunture complex of central Ethiopia, has yielded an assemblage and the partial mandible of a child (199-202). Based on

23 stratigraphic correlations with neighboring localities, Level E dates to ≈1.5 ma. The early human mandible (Gar IV E 0043) retains the right corpus with the dm1 and dm2 in occlusion, the developing I2, C1 and M1 evident in their exposed crypts, and the initial development of the P3 and P4 evident radiographically (202, 203; Fig. S23). Recent human dental development (145) provides an age-at-death of ≈3 years (see 202); given the apparently faster dental development of Early Pleistocene Homo (204), it was likely closer to ≈2 years of age. The teeth of the Garba IV immature mandible exhibit a number of irregularities associated with defective enamel development (202, 203, 205), including enamel pitting, vertical grooves on the buccal and lingual crowns, marked wrinkling of the occlusal surfaces of the unworn dm2 and M1, and reduced radio-opacity of the enamel (Fig. S23). In addition, the dm1 exhibits excessive occlusal wear for its limited time in occlusion, even for an Early Pleistocene child. As differentially diagnosed by Zilberman and colleagues (203, 205), the dentition exhibits amelogenesis imperfecta, which includes various combinations of incomplete maturation, calcification and mineralization of the deciduous and permanent teeth (206, 207). Amelogenesis imperfecta is produced by a number of genetic abnormalities with a diversity of inheritance patterns, and it may Figure S23. Distolingual view of the Garba IV- occur alone or as part of a syndrome (208). E43 immature partial mandible, with the fully Population incidences in recent humans have varied erupted and worn dm , the erupted and unworn 1 from ≈0.01% to ≈0.43%, with higher frequencies in dm2, and the M1 crown in its crypt. Reprinted from ref. 203. Copyright (2004), with permission from some local groups (207). Amelogenesis imperfecta is Elsevier. likely to have produced masticatory difficulties for the Garba IV child (205), especially in an Early Pleistocene Oldowan context, which may account in part for its young age-at-death. It is nonetheless unusual to find a case of one of these genetic disorders (despite the large number of mutations that can produce it) in the small sample of Early Pleistocene Homo remains, especially of very young individuals. Incidence: <1.0%

Lazaret 18 and 19 – Premolars Lazaret 18 and 19 (Fig. S24) consist of adjacent left mandibular premolars (12, 209), which derive from the late Middle Pleistocene (MIS 6) Middle Paleolithic Level C-II inf of the , in southeastern France (210, 211). The Lazaret 18 and 19 premolars are fully formed teeth, with occlusal wear and closure of the radicular apices. The Lazaret 18 P3 has a normal crown with buccal and lingual occlusal wear facets with moderate dentin exposure. Its root, however, has a distinct distal curvature, approximately one-third of the distance to the apex, rather than the usual modestly distal radicular deviation. The Lazaret 19 P4 has a misshapen crown and an exaggerated distal curvature of the root, resulting in a distal orientation of the occlusal surface relative to that of Lazaret 18 and the primary radicular of the two teeth. As a result of the distal tilting of its crown, the occlusal facet of Lazaret 19 is located on its mesial marginal ridge. Lazaret 18 exhibits normal mesial and distal interproximal facets, but the distal one makes contact with the mesial superior third of the root of Lazaret 19. The distal interproximal facet of Lazaret 19 is on its distal occlusal crown, for the absent M1. The Lazaret 19 crown exhibits a prominent central cusp with an associated reduction of the buccal and lingual cusps. The large central cusp is evident both occlusally and in the dentoenamel junction of the tooth. It represents a premolar odontome or dens evaginatus (212). The Lazaret 19 odontome conforms to type 5 of Schulge: “a tubercle arising from the occlusal surface obliterating the

24 central groove” (212: 2). They occur in recent human samples up to ≈4%, but are known principally among eastern Eurasians and Native Americans; they are very rare in western Eurasian and African samples (213, 214). They are otherwise undocumented among Pleistocene humans, although their presence has been inferred for the earlier Late Pleistocene Zhiren 3 (215; see below).

Figure S24. The Lazaret 18 and 19 mandibular left premolars in buccal and lingual views (B and L above), and CT imaging of the radicular structures in buccal, occlusal and lingual views (B, O and L below). In the buccal views, the Lazaret 18 P3 is left and the Lazaret 19 P4 is right. Images: M.-A. de Lumley and G. Becam, reprinted with permission from refs. 12 and 209. The more exceptional aspects of Lazaret 18/19 concern the radicular fusion of the two teeth, in combination with the distal root concavity and distal crown orientation of Lazaret 19. The crowns, although in contact in vivo as indicated by their interproximal facets, remained separate. The roots were joined along the full distal root of Lazaret 18 and the inferior three- quarters of the Lazaret 19 mesial root. In addition, the two pulp chambers were in communication along 5 mm of their middle portion. The pulp chamber and root canal of Lazaret 18 are otherwise normal, but those of Lazaret 19 exhibit multiple branches within the crown and the root apex. The developmental fusion of adjacent teeth is known clinically (and archeologically) (216-219), occurring in ≈0.5% of deciduous and ≈0.1% of permanent teeth. It is usually characterized by the fusion of adjacent crowns, often involving supernumerary teeth and occurring especially in the anterior dentition. The incidence of radicular fusion is not available, but it appears to be considerably less than the 0.1% reported for (mostly anterior) permanent teeth. Although a case of dens evaginatus and dental fusion has been reported (217), the fusion did not involve the premolars. The Lazaret 18/19 premolars therefore present two aspects (12), each of which is a rare occurrence in at least recent humans. They involve P4 but not P3 dens evaginatus, and radicular (rather than coronal) fusion with joined pulp chambers. It is possible that the deformation of the Lazaret 19 P4 and its associated fusion to Lazaret 18 are related to its dens evaginatus, through occlusal pressure. Yet, the normal lesion associated with dens evaginatus, and its clinical concern, is odontome fracture, pulpal exposure and radicular apical granuloma (212), and normally erupting premolars with odontomes are the rule rather than the exception (212, 214). It is therefore appropriate to consider the dens evaginatus and the radicular fusion of Lazaret 19 as separate abnormalities. Incidences: <5.0% and <0.01%

Malarnaud 1 – Incisors The Malarnaud 1 mandible, from southwestern France, was discovered in 1888 associated with Late Pleistocene fauna within a karstic system, but without archeological associations. Its morphology, and in particular the modestly retreating symphyseal profile with only a trace of a tuber symphyseos, led to its early association with the then known Neandertal mandibles (220-222). The associated fauna provides little paleoclimatic information, and the isolated mandible is therefore assumed to derive from the earlier Late Pleistocene, MIS 5 to 3b. It retains the corpus, the right ramus and most of the left ramus, including the right M1 in situ, the M3 crowns in their crypts, and all of its other alveoli (Fig. S25A). Although the bilateral canine to M2 alveoli are present and clearly identified as such, there are only two

25

(empty) alveoli mesial of the canine ones (Fig. S25B). As has been noted (e.g., 17, 222, 223), the individual experienced agenesis of two of its mandibular incisors. There is an insufficient number of symphyseal (or intercanine) alveoli, and there are no unerupted teeth within the corpus (222; Fig. S25D). The Malarnaud 1 mandible represents an early adolescent, with full eruption of all of its teeth except the M3s, as indicated by alveoli indicating post-mortem dental loss and a distal interproximal facet on the M1. The M3s are barely at complete crown formation, providing a recent human age-at-death of 14- 15 years (145); given precocious M3 development in Late Pleistocene humans (8, 224), a modestly younger age is indicated. There is variation as to which of the mandibular incisors have been inferred to have failed to form, the I1s (222) or the I2s (17). Which incisors are absent is indeed ambiguous. There is a large alveolar diastema ≈5 mm wide just to the left of the symphyseal midline, such that the remaining incisor alveoli are adjacent to the midline on the right and distally displaced on the left. The right one has a vertically set tapering socket, whereas the left one is larger and curves mesially to its apex Figure S25. The Malarnaud 1 adolescent mandible. A: (see radiographs in 222). It is therefore likely occlusal view of the dental arcade; the arrow is at the that the right I2 is absent, whereas it is the left midline, as indicated by the mentum osseum morphology I that is absent. and the genioglossal muscle markings. B: occlusal detail 1 of the anterior alveoli, including the bilateral canine and Dental agenesis is relatively rare in the two incisor alveoli. C: anterior/labial view of the Pleistocene humans. Third molar agenesis, canine and incisor alveoli and the mandibular symphysis. although common among recent humans, occurs in only a few cases prior to the Late Upper Paleolithic (17), being present in the Middle Pleistocene Chenjiawo 1 and the Late Pleistocene Liujiang 1 and Dolní Věstonice 16 (67, 225), as well as the early hominin Omo 75-14 (226). No other cases of pre-MIS 2 incisor agenesis have been documented. In a pooled global sample of recent humans (n = 48,274), I1 agenesis occurred in 0.3%, and I2 agenesis in 0.2% (227). The probability of agenesis for either both I1s or both I2s would be about half of these percentages, given about equal occurrence of unilateral versus bilateral I1 agenesis (227). It is unclear what the probability would be for bilateral agenesis of one of each incisor, but it would be distinctly <0.1%. Incidence: <0.1%

Oase 1 and 2, Denisova 4 and 8, and El Haroura 1 - Distal Molars The sites of Peştera cu Oase in southwestern , in the Altai Mountains of western Siberia, and El Haroura I in Morocco have yielded various Late Pleistocene human remains. The first are Early Upper Paleolithic in age (MIS 3) modern humans (228, 229), the second are Middle Paleolithic associated (MIS 3) archaic humans (230-233), and the third is a (Aterian) early Late Pleistocene non-modern human from northwest Africa (234). Although they are spread from the Atlantic to Siberia and have diverse morphological (and paleogenetic) affinities, they share exceptionally large distal molars (Fig. S26). All of these remains are, or include, third molars, maxillary ones for Oase 2 and Denisova 4 and 8 and mandibular ones for Oase 1 and El Haroura 1. The isolated teeth from Denisova are attributed to M3s based on a combination of occlusal morphology, their degrees of occlusal wear and the absence of distal interproximal facets; the others are securely in their alveoli.

26

If the crown diameters of the Denisova 4 and 8 M3s are used to create an “area” (length x breadth), the resultant values of 192.6 and 209.5 mm2 respectively far exceed those for almost all of the known Middle and Late Pleistocene human maxillary M3s (Fig. S26A). They are respectively 4.13 and 5.00 standard deviations from a pooled Late Pleistocene archaic human sample (p = 0.0003 and p = <0.0001). The one value close to them is the in situ (unerupted) Oase 2 left M3 (186.3 mm2) (235), which is 3.63 standard deviations (p = 0.0017) from a pooled Early/Mid Upper Paleolithic mean (the right M3 is similar in size, but does not provide a length). All three of the M3s are completely outside the range of Middle and Late Pleistocene human M3s, and they are matched only by those of Austalopithecus and approached by a few initial Pleistocene Homo specimens (230, 235, 236). To these human remains with megadont molars can possibly be added the Middle Paleolithic (MIS 3) central Asian late archaic human remains (with Neandertal affinities) from Obi- Rakhmat (237, 238). The partial M3 of Obi- Rakhmat 1 has a buccolingual crown breadth of 14.6 mm, which is close to those of 14.6 and 14.9 mm for Oase 2 and 14.7 mm for 3 Figure S26. Comparisons of maxillary (left) and mandibular (right) third molar both Denisova M s. crown “areas” (length x breadth) in mm2 for Middle and Late Pleistocene However, an isolated 3 3 samples, the M s of Oase 2 and Denisova 4 and 8, and the M3s of El Haroura 1 Neandertal M from La and Oase 1. Mid Middle Pleistocene: ≈600 – ≈300 ka BP; Late Middle Chapelle-aux-Saints in Pleistocene: ≈300 – ≈130 ka BP; Late Pleistocene Archaic: ≈130 – ≈40 ka BP; western has a Middle Paleolithic modern humans (MPMH): ≈90 ka BP; Early and Mid Upper buccolingual breadth of Paleolithic ≈40 – ≈24 ka BP; Late Upper Paleolithic: ≈24 – ≈12 ka BP. The 15 mm but a mesiodistal samples sizes for the sequential samples for the maxillary and mandibular M3s one of only 9.3 mm, respectively are: 24 and 25, 13 and 13, 32 and 42, 7 and 9, 28 and 21, and 17 and providing it with an 14. unexceptional “area” of 139.5 mm2 (see Fig. S26A). It is therefore unclear whether the Obi-Rakhmat 1 M3 exhibited the distal molar megadontia of the Denisova and Oase M3s. In addition to the exceptionally large M3s of these (otherwise unrelated) Late Pleistocene humans, the Oase 1 and El Haroura 1 mandibles contain unusually large M3s (Fig. S26B). Their respective right/left average “areas” of 170.5 and 171.5 mm2 are 2.74 and 2.83 standard deviations above the Early/Mid Upper Paleolithic and late archaic human means respectively (p = 0.008 for both). They are unmatched by any of the known late Middle to Late Pleistocene human M3s, and it is only in the early to mid Middle Pleistocene that their values are approached by those of Arago 13 (165.7 mm2), Tighenif 2 2 2 (167.5 mm ) and B-II (162.8 mm ). The large Oase 1 M3s parallel the large maxillary ones 3 2 of Oase 2, but the large M3s of El Haroura 1 are not associated with large M s (114.0 and 114.4 mm ) in the otherwise rather megadont northwestern African Aterian sample (239). The other Aterian M3s (132.8, 135.5, 141.5 and 155.5 mm2) are among the larger of the late archaic ones; yet the largest of these other Aterian M3 “areas” (Dar es-Soltane 4) is nonetheless matched by those of the Krapina 8, 5 and 9, and Shanidar 6 Neandertals (150.8 – 156.2 mm2) and the Pataud 1, Předmostí 3, 2 and Qafzeh 9 M3s (151.7 – 152.3 mm ) (Fig. S26B).

27

It is not clear whether these very large M3s from Denisova Cave, Peştera cu Oase and El Haroura I merely represent megadont isolated populations in southwest Siberia, southeast Europe and northwest Africa during MIS 3, or whether there is some other process involved. The M3s are aberrant in their size, if not their morphology, for Middle or Late Pleistocene humans and the M3s are exceptional for later Middle and Late Pleistocene humans. To the degrees that Denisova 4 and 8 may be related, and the same for Oase 1 and 2, their distal molar megadontia may be populational rather than individual phenomena, and the same may apply to the El Haroura 1 in an Aterian context. Yet, there is no reason to infer any particularly close relationship between these three samples other than being MIS 3 humans. Given the relatively large samples available especially for Late Pleistocene distal molars (Fig. S26), these individuals/samples (especially for the M3s) are exceptional. Probabilities: <0.01% Denisova 8, <0.1 Denisova 4, <1.0 El Haroura 1 and Oase 1 & 2

Pataud 1 - Paramolars The late Mid Upper Paleolithic (Late Gravettian/Proto-Magdalenian; MIS 3) level of the in southwestern France (240, 241) has yielded the partial skeletons of six individuals, three adults, two infants and a juvenile (242, 243). The remains appear to have been manipulated post-mortem, given their dispersion in the site and the conspicuous absence of certain elements (244). One of these individuals, the young adult Pataud 1, pelvically and cranially sexed as female, presents an unusual pattern of supernumerary teeth (28, 245, 246).

Figure S27. The right maxilla of Pataud 1 in lateral (above) and occlusal (below) views; the occlusal view has the more recently excavated teeth, not all of which fit entirely within their alveoli due to shrinkage of the alveoli while empty. One of the supernumerary teeth (paramolars) is in situ adjacent to the M1 and especially the M2. The other one is in place in the occlusal view and indicated by the arrow, and it is separate in the upper right in the buccal view. Occlusal and separate paramolar images courtesy of Sébastien Villotte (Université de Bordeaux, Pessac, France). The Pataud 1 dentition has fully formed permanent teeth with minimal to moderate degrees of occlusal attrition on all but the M3s. The M3s are fully formed and erupted, but the left M3 is distally oriented and slightly short of the M1-M2 occlusal plane, and the right M3 is impacted against the distal surface of the M2 (Fig. S27). There is widespread evidence of periodontal inflammation, as well as large mandibular retromolar alveolar voids. In combination with these oral health issues, there are two supernumerary teeth, adjacent to the right M2 and M1/M2 (28; Fig. S27). The more lingual one is obliquely set, with its root in the buccal M2 alveolar bone and fused to the buccal M2 root, and with the crown against the buccal M1 and especially M2 crowns. Its crown sits largely within the buccal interproximal space of the M1 and M2. The second supernumerary tooth is now separate but sets into the resorptive space above the M1 and the other supernumerary tooth. The more lingual supernumerary tooth is conical, but the more buccal one has a larger cusp and a small one separated by a fissure, along with a bifurcated root. Supernumerary teeth in the molar region are among the least common of such supplemental teeth (190-194), occurring in <0.1% to 0.8% of recent clinical samples. They occur mostly distal to the M3s (distomolars), and small ones adjacent to the regular molars (paramolars) are noted primarily as

28 individual cases (196, 247). In two recent human samples, paramolars had frequencies of 0.06% and ≈0.2% (192, 247). They may be lingual or buccal and can be maxillary or mandibular; mandibular versus maxillary predominance is variably reported (196, 197, 248, 249). Paramolar crowns can be simple conical shapes or more molariform (250). They are rarely bilateral (196, 251), but double unilateral ones similar to those of Pataud 1 appear to be unreported. Paramolars are usually located adjacent to the M2/M3, and rarely occur at the M1/M2 (247). As with supernumerary teeth in general (190, 192), they are more common in males. There is little clarity on their non-syndromic etiology (192). The already low probability of finding a paramolar in an isolated Pleistocene specimen, even with the relatively abundant sample of Upper Paleolithic remains, would be further reduced by it being double, at M2 to M1/M2, and in a female. Incidence: <0.01%

Pataud 6 – Premolar In addition to the supernumerary paramolars of the late Gravettian (MIS 3) Pataud 1, the stratigraphically similar early juvenile partial skeleton of Pataud 6 (243) exhibits an abnormal presentation of one of its developing premolars (28, 245, 246). The Pataud 6 right maxilla (Fig. S28) retains the deciduous molars (dm1 and dm2) fully formed and in occlusion, with the root apices closed. The developing crowns of the I2, C1, P3 and P4, at appropriate relative stages of calcification, are present in their crypts. The first three permanent teeth are normal and in the appropriate positions within the maxilla. The P4, however, although it was forming normally, is inverted in its crypt. Given that it is nested within the roots of the dm2, its inverted position is not an artifact of post-decomposition movement; it is in its in vivo position. Developmental inversion may affect any of the permanent teeth, as well as deciduous and supernumerary teeth (252, 253). They are extremely rare (less than a few dozen reported cases globally), are exclusively reported as individual cases, and appear to most commonly involve impacted third molars. They Figure S28. Buccolingual radiograph of the Pataud 6 right maxilla. The rarely include premolars (254, dm1 and dm2 are in occlusion with closed root apices. The developing 255). The presence of an inverted crowns of the I2, C1, P3 and P4 are in their crypts, those of the P3 and P4 4 P in Pataud 6 is therefore truly nestled within the roots of the dm1 and dm2 respectively. The arrow unusual. indicates the inverted crown of the P4. Radiograph courtesy of Robert L. Tompkins (photographer). Incidence: <0.01%

Qafzeh 15 – Premolar The Qafzeh 15 partial skeleton derives from an early Late Pleistocene (MIS 5) Middle Paleolithic context in northern Israel (5, 79, 80), and it retains substantial portions of the cranium and mandible, the developing dentition, pieces of the right scapula and both humeri, plus crushed and cemented vertebral and costal remains (5). The dentition, with the deciduous canines and molars plus the permanent incisors and first molars in occlusion, indicates an age-at-death ≈ 9 years. Each maxilla retains the developing 1 3 4 2 crowns of the C , P , P and M . The right mandibular corpus has the developing crowns of the C1, P3, and M2, plus the M3 crypt. The region of the right P4 was broken post-mortem, and the P4 (if present) was lost.

29

The left corpus also retains the crowns of the C1, P3 and M2, as well as the M3 crypt. However, as noted in detail (5, 223), the intact left corpus lacks any evidence of the developing P4 crown (Fig. S29B). The region of the P4, just distal of the P3 and below the dm2, presents normal mandibular bone with no trace of a developing crown. There is no damage to the left corpus in the vicinity of the deciduous molars and hence the developing premolars (Fig. S29A). Therefore, the Qafzeh 15 late Figure S29. Lateral view of the Qafzeh 15 mandible (cast) (A), and juvenile exhibits agenesis of the left buccolingual radiograph of the left P4 (223). It is not possible to mandibular corpus of Qafzeh 15, with determine whether the right P4 was the developing P3 (arrow) (B). The left developing and lost post-mortem or I1, I2, dc1, dm1, dm2, and M1 are apparent also experienced agenesis. The right in occlusion, and the developing crowns and left maxillae retain all of the of the P3 and M2 are evident in the expected occlusal and developing radiograph. Note the absence of a permanent teeth. developing P4 crown distal of the P3 crown. Radiograph reprinted with Agenesis of the mandibular permission from ref. 223. second premolar (P4) is the most common location in the dental arcade for agenesis, after the third molar. In a large pooled recent human sample (n = 48,274), it occurred in 3.1% of individuals (227). Unilateral, versus bilateral, P4 agenesis occurs in ≈54.5% of cases with P4 agenesis (227), such that the incidence would be 1.7% for unilateral agenesis versus 1.4% for bilateral absence of the P4s. Incidence: <5.0%

El Sidrón Adult 2 and Adolescent 3 – Canines The late Middle Paleolithic (MIS 3) site of El Sidrón in northwestern Spain has yielded >2,500 remains of seven adult and six immature individuals, which derive from a closely related Neandertal group (256-258). Although the human remains were disassociated, had been humanly processed, and were displaced within the karstic system, it has been possible to associate most of the remains by individual. Two of these individuals, the El Sidrón Adult 2 and probably the El Sidrón Adol. 3, exhibit anomalies of dental formation/eruption and associated degenerative lesions (259).

Figure S30. The El Sidrón 2 left mandible (SDR 008), in medial view (A) and mediolateral radiograph (B). The lower left arrow on the photograph and the lines on the radiograph indicate the C1, which formed within the corpus and failed to migrate mesially with mandibular development. C: mandibular cross-section at the mesial M1, showing the enlarged inferior alveolar nerve canal and the lateral opening around the mesial root of the M1. D: mandibular cross-section at the mesial M3 and just above the the cervix of the left C1. Images: A. Rosas. Reprinted from ref. 259. Copyright (2013), with permission from Elsevier. The primary abnormalities of the El Sidrón Adult 2 mandible (259) involve an abnormally formed and retained left deciduous canine and the unerupted position of the left permanent canine within the mandibular corpus

30

(Fig. S30). The dc1 is peg-like and has a crown extension, but it was fully erupted between the I2 and the P3 with normal occlusal wear. The completely developed C1 lies horizontally within the corpus below the M2 to distal of the M3, such that the occlusal edge is below the M2 mesial root and the radicular apex is ≈3 mm distal of the M3 distal root. The C1 is lingual of the inferior alveolar nerve canal (Fig. S30D). It is close to the lingual mandibular surface, inferior of the mylohyoid line, and the root is visible through a medial fenestration of the mandibular corpus (Fig. S30A).

Associated with this distal retention of the C1, there is an osseous void (or cyst cavity) in the area around the C1, extending from it and the M2 root mesially to the I2 (Figs. S30B and S30C). The walls of the cavity appear to have been remodeled multiple times, but the roots of the teeth extending into the cavity appear unaffected and do not exhibit resorption. The enlarged cavity opened buccally around the mesial M1 to P4 roots and the I2 root. The maxillary dentition of El Sidrón Adol. 3 (259) retains the I1 to P4 plus the M2 and M3. All of these teeth show minimal to moderate, and age appropriate, occlusal wear except the unerupted M3. The mandibular dentition preserves both P4s, the left P3 and the retained left dc1. Associated with these teeth (given shared patterns of dental enamel hypoplasias) is an unerupted left mandibular canine. Given the full eruption and occlusal wear of all of the other permanent teeth except the unerupted M3, plus the retained dc1, the left C1 was either abnormally erupted such that it did not occlude with the other teeth or it was retained within the largely absent mandibular corpus. Following Dean and colleagues (259), the primary developmental abnormality in El Sidrón Adult 2 was most likely the developmental malformation of the dc1. Although the dc1 erupted into normal occlusion, it may have affected the development and especially the mesial migration of the C1 during longitudinal mandibular growth, thereby resulting in its position below the distal molars. The abnormal cavity within the mandibular corpus and its buccal openings are likely secondary effects of the retained permanent canine and subsequent infection spreading through the mandibular corpus. The lack of wear on the El Sidrón Adol. 3 C1, along with the retained dc1, may indicate a similar dental developmental anomaly. If these conditions have a genetic basis, it is interesting that these two individuals share at least the same mtDNA haplotype (256; see discussion in 259). The probability among recent humans of this malformation is unknown, but it appears to be unique to the El Sidrón Adult 2 and Adol. 3 among Pleistocene humans (see also discussion of the Subalyuk 2 deciduous canine below). Unknown incidence

Subalyuk 2 – Deciduous Canine The Middle Paleolithic site of Subalyuk (MIS 3-4), in northern (260), has yielded the partial remains of a younger adult (Subalyuk 1) and of a child (Subalyuk 2) (136, 261, 262). The latter specimen retains a largely complete neurocranium, the separate maxillae with most of its developing dentition, four vertebral centra and a tibial diaphysis. The child’s remains do not present any pathological

Figure S31. The Subalyuk 2 immature maxilla in occlusal view with the geminated left dc1 indicated (cast; left) and the left dc1 in labial view (right). Right image courtesy of Robert G. Franciscus (University of Iowa, Iowa City, IA). lesions beyond a trivial I1 linear hypoplasia (136), and the mixed maxillary dentition is largely present (262). In the dentition, the right di1 to dm2 and the left dc1 to dm2 are in occlusion (Fig. S31). Developing and unerupted permanent teeth consist of the right I1 and I2,

31 plus both M1s. It is dentally aged to ≈3 years (145, 262). The right dc1 is normal, as are the right deciduous incisors and the deciduous molars. However, as described by Bruszt (263) and Thoma (262), the left dc1 is “an anomalous tooth having no comparative significance,” merely a “rare schizoid variation” (262). The tooth presents a double crown, with a smaller, pointed crown fused to the mesial side of the full normal crown (Fig. S31). It has a fusion line, or sulcus, between the two portions of the crown labially and lingually. Labially the sulcus continues along the root to the apex, and where exposed near the cervix it continues lingually. The root otherwise appears largely conical and single. The pulp chamber is single in the root, but it becomes double within the crown. The distal portion of the left dc1 crown is generally similar to the normal right dc1, with minor differences (262), and the mesial portion is both more prominent occlusally and asymmetrical. The extra mesial portion on the dc1 is not the fused on di2 (although the left di2 is not preserved), because the normal di1 and di2 alveoli are present in the left maxilla. It therefore represents an additional portion of the dc1, apparently deriving from the same tooth germ (264, 265). The combination of a largely (but not entirely, given the labial and lingual sulci) single root with a double crown agrees best with a partial separation of the original dc1 tooth germ into two teeth, or a gemination of the tooth (also known as schizodontia) (264). It does not appear to represent fusion of the dc1 with a supernumerary tooth, given in particular the shared root pulp cavity. Cases of fusion with a supernumerary tooth and gemination are often combined as “double teeth.” In recent human samples of deciduous teeth, frequencies of “double teeth” vary from 0.1% to 1.5% with an overall frequency of ≈0.5%, of which a distinct minority are diagnosed as gemination (265-267). In addition, the majority of the deciduous double teeth (including geminations) are mandibular and incisors (265-267). The frequency of geminated maxillary deciduous canines is not available, but it is evident that it is a markedly rare dental developmental variant among recent humans. It would be exceptional to find one in the two dozen Late Pleistocene archaic human dc1s known. Incidence: <0.1%

Sunghir 2 – Lateral Incisor The Early/Mid Upper Paleolithic (MIS 3) Sunghir 2 skeleton, of an early adolescent male from northern Russia (8, 98-100; see above), presents an unusual lack of interproximal and occlusal wear on his dentition, for his age and context (see above; Fig. S8). In addition, his anterior maxillary dentition presents a small and erupted supernumerary tooth (Fig. S32). On the mesiolingual corner of the Sunghir 2 right I2 is a small, largely conical tooth, with a ≈1.8 mm mesiodistal crown diameter. The occlusal edge of the supplementary tooth is against the mesiolingual edge of the I2, at the lingual Figure S32. The anterior dentition of end of the (very small) mesial Sunghir 2 with the small supernumerary marginal ridge and slightly tooth adjacent to the right I2 indicated, in mesial of the lingual tubercle. occlusal (above) and oblique lingual It was previously (8) (below) views. described as a possible enamel pearl, but it has a distinct cylindrical root along side that of the I2. It is therefore best seen as a small supernumerary tooth, which became fused to the mesiolingual crown and root of the I2. Supernumerary teeth in the permanent dentitions occur in recent human populations in 0.1% - 3.8% of individuals (189-192), but frequencies for them adjacent to maxillary incisors vary from absent to

32

≈0.7% (190-193). Non-descript conical ones are the most common form (192, 193). In addition, supernumerary teeth are consistently more common in males (190, 192). The developmental fusion of adjacent teeth is known clinically (and archeologically) (216-219), occurring in ≈0.5% of deciduous and ≈0.1% of permanent teeth. It is usually characterized by the fusion of adjacent crowns, often involving supernumerary teeth and occurring especially in the anterior dentition. The presence of a small, conical supernumerary tooth in the maxillary incisor area of Sunghir 2 is therefore unusual but not exceptional. It is unclear how the partial fusion of it to the I2 would affect the probability of finding this arrangement in the anterior dentition of Sunghir 2. Incidence: <1.0%

Zhiren 3 – Incisors and Premolars The edentulous Zhiren 3 anterior mandibular corpus, from the early Late Pleistocene (MIS 5) of Zhirendong () in southern China, retains the alveoli from the right P4 to the left P4 (268). The mandible and the Zhiren 1 and 2 isolated teeth lack an archeological context, but stratigraphically associated mammalian remains, U-series dates and paleomagnetism (268, 269) provide a secure context. The fragmentary alveoli attached to the Zhiren 1 M3 indicate advanced periodontal inflammation and resorption, and the Zhiren 2 M2 or M3 has a large distolingual cervical carious lesion (215). However, it is the abnormalities of the Zhiren 3 mandible that appear to have a developmental origin (215). The well-preserved anterior alveoli of Zhiren 3 present labiomesial rotation of both central incisors (Fig. S33), normally characterized as the “winging” of the teeth (270). Although apparently genetically determined (271) and generally more frequent among eastern Eurasian (and derived) populations, incisor “winging” is unexceptional among recent humans (272). Zhiren 3, however, is apparently the only known Pleistocene case of mandibular incisor “winging”. The Zhiren 3 mandible also exhibits marked enlargement of both P3 alveoli (Fig. S33), which appears to have occurred without ante-mortem loss of the teeth. The superior half of the right P3 socket and the superior two-thirds of the left one conform to the shapes of the premolar roots, although they are modestly porous and lack the lamina dura. Below each superiorly altered socket, the apical regions are expanded and open buccally on to the lateral mandibular corpus. The left socket in particular rounds out on to the lateral corpus, although a porous edge marks its inferolateral margin. The preserved alveoli of the adjacent teeth on both sides (the C1s and the left P4 in particular) do not show any degeneration Figure S33. The Zhirendong 3 anterior beyond trivial porousness of the left C1 alveolar crest. mandibular corpus in anterior (above), occlusal The Zhiren 3 mandible therefore exhibits bilateral P3 (below) and left lateral (right) views. Scale: 5 granulomata without any involvement of the adjacent and cm. Images courtesy of Xiujie Wu (Institute of more mesial teeth. Vertebrate Paleontology and Paleoanthropology, Beijing, China). The bilateral presence of P3 lesions in the absence of other pathological changes in the mandible suggest an etiology other than periodontal or a systemic disorder affecting the alveolar process (215). A probable etiology is dens evaginatus, in which there would have been a bilateral P3 occlusal tubercles (212, 273; see Lazaret 19 above). These tubercles extend well above the occlusal plane, are frequently subject to breakage from occlusal forces, and (as they usually contain pulpal extensions) can be subjected to pulpal and periapical infections and subsequent necrosis. In recent human samples they occur up to ≈4%, but are known principally among

33 eastern Eurasians and Native Americans; they are very rare in western Eurasian and African samples (213, 214). In the absence of the teeth, the diagnosis of dens evaginatus in Zhiren 3 is not definite. However, the restriction of alveolar degeneration to the P3s and the advanced states of those changes strongly indicate an unusual etiology. The presence of incisor “winging” and the probable occurrence of dens evaginatus in Zhiren 3, both of which occur in higher frequencies among eastern Old World populations, could suggest a relationship between the Zhirendong human remains and recent eastern populations. Yet, given the ≈100,000 year gap between Zhiren 3 and those recent humans, and the complex population processes since then (274), there is no necessary connection of Zhiren 3 to recent eastern Eurasians. It geographical location should therefore not influence assessments of its likely incidence. Incidence: <5.0%

Vertebral Abnormalities Arene Candide 2 – The Late Upper Paleolithic (final Epigravettian, MIS 2) site of Arene Candide in northwestern Italy has yielded a ≈20 individuals, two-thirds of which were primary burials (32, 44, 45). Six sacra of the adult skeletons present variations in terms of the sacralization of the fifth (Arene Candide 10) and the first coccygeal one (Arene Candide 3) and in the cranial extents of the sacral hiatus (45). These variations are relatively common, but the sacrum of Arene Candide 2 presents additional developmental variations (Fig. S34).

Figure S34. Ventral (A), dorsocranial (B), oblique left dorsolateral (C) and lateral left (D) views of the Arene Candide 2 sacrum. Note the partial age-related fusion of the ventral sacral bodies, the anomalous non-fusion of the S1 neural arch to the (albeit damaged) neural arch of the S2, and the partial fusion of the left S3/S4 and S4/S5 synchondroses. Images courtesy of Vitale S. Sparacello (Université de Bordeaux, Pessac, France) (copyright MIBAC- SABAP Liguria).

The Arene Candide 2 skeleton is of a young adult male based pelvic morphology, pubic symphysis metamorphosis, and the proximal clavicular epiphysis (275, 276), buried with the remains of Arene Candide 3 and 4 arranged around him (32, 276; see Arene Candide 3 above). The minimal degree of ventral sacral body fusion (Fig. S34A) provides a similarly young adult age-at-death (277), and its ventral body fusion is normal in having the S1/S2 bodies well separated but only an open fusion line evident between each pair of the more caudal bodies. Dorsally, the neural arches of S1 to S5 are complete and those of S2 and S3 are fused together. However, the neural arch of S1 is separate from that of the S2 (Figs. S34B to S34D). The articular facets are present on the caudal arch of the S1, and the right one articulates with the cranial facet of the S2; the corresponding portion of the left S2 was lost post-mortem. The laminae and spinous processes of S1 and S2 are separate. There is also incomplete fusion of the S3/S4 and S4/S5 neural arches, principally on the articular facets. There is the remains of the left S2/S3 lateral portion fusion line ventrally at the caudal auricular surface (Figs. S34A and S34D), and the S3/S4 lateral portions are only partially fused dorsally

34 and laterally (Fig. S34B and S34D). The etiology of these inconsistent fusions of normal sacral synchondroses is unclear; it is also uncertain whether it might be related to the bilateral absence of lesser trochanters from the Arene Candide 2 femora (276; see below) or the presence of only four for the individual (32). Unknown etiology

Atapuerca-SH Pelvis 1 – L5 The mid Middle Pleistocene (Acheulian; MIS 11–12 ) site of Sima de los Huesos in the Sierra de Atapuerca (Atapuerca-SH), north-central Spain has yielded thousands of human remains from about two dozen individuals (115, 116). The remains reached the Sima de los Huesos through a long vertical shaft within the karstic system, and became fragmented and mixed in the process. It has nonetheless been possible to reassemble and reassociate a largely complete male pelvis (Atapuerca-SH pelvis 1) with its five lumbar vertebrae (278). Based on pelvic morphology and the metamorphosis of the auricular surfaces, the Atapuerca-SH pelvis 1 derives from an older male. The lumbar vertebrae and sacrum exhibit a number of degenerative changes, including ventral new bone formations, exaggerated body wedging, a Schmorl’s node, and multiple asymmetries, indicating osteoarthritis, kyphosis and (278). In addition to these degenerative changes, the fifth lumbar vertebra presents an incomplete ossification of its left neural arch, or , associated with a right deviation of the spinous process (Fig. S35). The absence of a synchondrosis surface on the left side of the spinous process and of post-mortem breakage indicates that the left lamina was developmentally absent. Partial or complete separation of the neural arch, unilaterally or bilaterally, is moderately common among recent Figure S35. Cranial view of humans, having been documented skeletally in 4.5% (n = 7,063) of a the fifth lumbar vertebra (L5) pooled recent human adult sample (279-282), although substantially of Atapuerca-SH Pelvis 1 and higher frequencies have been noted in prehistoric Native American left lateral view of its spinous samples (283). In the largest individual recent human sample (n = process (below left). Note the 4,200), 85.2% of the separations (n = 183) were in the L5, and of developmental gap between the those, 17.3% (n = 27) involved unilateral separations (or 0.6% left pedical and articular facets unilateral L5 occurrence in the sample) (280). Moreover, it is and the base of the spinous unclear to what extent neural arch separation in the sample involves process (absence of the left the pedicles as opposed to the laminae. lamina) and the absence of evidence of fracture or There appear to be congenital predispositions to develop synchondrosis on the spinous neural arch separations (284, 285). Yet, in recent humans, the process. Modified from ref. incidence tends to increase with age, especially among athletes (281, 278. 282, 285), indicating a role for elevated levels of stress in the lumbar region in spondylolysis presence. Although serious trauma could be involved (284), the absence of clear traumatic lesions in the Atapuerca-SH L5 suggests a developmental and persistent stress etiology for the absence of the left L5 lamina. The asymmetries in the vertebrae (including the sacrum) (278) may related to the spondylolysis, and it may be related to the evidence for spondylolisthesis of the L5 on the sacrum, since the former often predisposes the individual for the latter (285). The presence of a spondylolysis in the Atapuerca-SH Pelvis 1 L5 is unusual, especially given the minimal number of Middle Pleistocene lumbar vertebrae (116) despite its moderately common occurrence

35 among recent humans. However, its unilateral manifestation is rare, at least relative to recent human mature samples. Incidence: <1.0%

Grotte-des-Enfants 1 – T5 The Grotte-des-Enfants (Grotta dei Fanciulli) in the (Baoussé-Roussé) cliff of northwestern Italy yielded a long Late Pleistocene sequence (286-288), including several Late Upper Paleolithic (Epigravettian; MIS 2) layers. Directly dated to the final Epigravettian are two infants (Grotte- des-Enfants 1 and 2), that were probably an intrusive burial from Layer B into Layer C (288). The Grotte- des-Enfants 1 child, dentally aged to ≈3 years, retains a remarkably complete skeleton, including a that lacks only the S5. The Grotte-des-Enfants 1 skeleton presents a series of minor areas of porosity and/or new bone formation, as well as minor non-specific stress indicators (6). In addition, it has an anomalous articular facet on the fifth thoracic vertebra. The T5 retains most of its neural arch with the base of the (now broken) spinous process. At the intersection of the laminae dorsally, on the cranial surface, there is a modestly elevated amygdaloidal surface, which represents a supplementary articular facet (6). The adjacent caudal surface of the T4 is preserved, and it does not have a matching facet (nor do any of the other thoracic neural arches present anomalous facets, to the extents preserved). It is not clear why there would be a one-sided additional articulation in this location. Unknown etiology

Kebara 2 – Sacrum and Lumbar Vertebrae The Late Pleistocene (MIS 3) Middle Paleolithic levels of in northern Israel have yielded two burials with the partial skeletons of an infant (Kebara 1) and a young adult (mid-third decade) male (Kebara 2) and the fragmentary remains of additional late archaic humans (289-291; see 292). The latter partial skeleton retains the mandible, complete upper limbs, the pelvis and a remarkably complete vertebral and costal skeleton (290). The Kebara 2 skeleton presents a variety of minor osteoarthritic and traumatic lesions (4,293), but it is its abnormal development and synchondrosis fusion of the lumbar and sacral vertebrae that are of concern. The largely complete sacrum of Kebara 2 appears generally normal in overall morphology (despite some post-mortem distortion) (294), and the lumbar vertebrae are generally as expected morphologically (293). However, the caudal three lumbar vertebrae (L2 to L5) and the second sacral one (S2) lack normal bony formation of their spinous processes, the S1 lacks its neural arch dorsal of its zygapophyses, and the sacrum exhibits an unusual and asymmetrical pattern of partial fusion (4,293; Fig. S36).

36

Figure S36. Ventrodorsal radiograph of the Kebara 2 sacrum (left), ventral view of the Kebara 2 sacrum (right), and cranial view of the Kebara 2 L5 (below center), showing the incomplete and irregular ossification of the sacral elements and the L5 spinous process. Images courtesy of Charles E. Hilton (University of North Carolina at Chapel Hill, Chapel Hill, NC). Radiograph courtesy of The laminae of L2 to L5 terminate dorsally in normal midline Henri Duday (Université de fusions, but they have only irregular rounded knobs for the bases of the Bordeaux, Bordeaux, France) spinous processes. This is particularly evident on the L5 (Fig. S36), and Baruch Arensburg (Tel which has no presence of a spinous process. The L2 to L4 have slightly Aviv University, Tel Aviv, more, but minimal, development of the base of the process. This Israel). agenesis of the L2 to L5 spinous processes is continued with an absence of the dorsal neural arch of the S1 and a very reduced dorsal midline of the S2. The spinous process of the L1, and the more cranial vertebrae, are developmentally normal. On the sacrum, the S1 is completely unfused to the S2, including between the bodies and the articular facets and through the lateral portions (Fig. S36). Non-fusion (or partial fusion) of the S1/S2 bodies is common in young adults (277), but the non-fusion of the S1/S2 lateral portions and articular facets is not normal post-adolescence (295). In addition, although the fusion lines between the S2/S3 and S4/S5 bodies are basically closed (Fig. S36), the bodies of the S3 and S4 are distinctly separate. The fusion of the S2/S3 and S4/S5 bodies is normal by the mid-third decade, but complete non-fusion of the S3/S4 bodies is not (277). The left side also has a partial fusion of the S3/S4 lateral portions (the right side is too damaged to assess), a developmental delay which is probably related to the sacral body non- fusion. There are also bilateral ventrolateral extensions of bone, along the S1 auricular surfaces, fused to the S2 lateral portions. They should represent portions of the sacroiliac epiphyses, which had fused to the S2 but not to the S1, although they are considerably more massive than the usual sacroiliac epiphyseal plates. They should have been fused (or nearly fused) by the age-at-death of Kebara 2 (295). These auricular epiphyses and their non-fusions may be related to the marked asymmetry of the auricular surfaces. The non-fusion of the S1 to the S2 may have produced a minor spondylolithesis of the S1 on the S2. In combination with the agenesis of the lumbar spinous processes, it may have engendered postural instabilities. In addition to these developmental deficiencies, there are other abnormalities of the vertebrae, some of which may be developmentally or secondarily related to these changes (4). The vertebral canals of the L4 and L5, and especially the latter, are constricted caudally. There are enlarged and supplemental vascular canals in the T11 and T12. The cranial L2 body has a minor Schmorl’s node. The spinous process of the T5 sustained a fracture. There are minor osteophytes on the vertebral bodies of the T10 and L1 to L4. There are modifications of the T3 to T9 costotransverse articular facets. There are osteoarthritic changes to the atlantoodontoid and C2/C3 plus L5/S1 articular facets. A ventral bony extension into the lateral portion of the right first sacral vertebral foramen reduced the foraminal diameter. In addition, the L1 presents lumbar (considered below with the Shanidar 3 L1 facets).

37

It is therefore apparent that the lumbar and sacral vertebrae of Kebara 2 present a series of ossification abnormalities, of which only the open S1/S2 bodies can be seen merely as a modest delay of a normal developmental sequence. Although there is middle thoracic evidence of trauma, in the T5 spinous process, these changes are not traumatically related. The various degenerative changes in the vertebrae may be secondarily related to the developmental abnormalities, but they are not etiologically primary. Yet, as emphasized by Duday and Arensburg (4:192), the combination of ossification deficiencies is “frankly abnormal.” Unknown etiology

Nariokotome (KNM-WT 15000) – Craniocervical The Nariokotome (KNM-WT 15000) largely complete skeleton of the 1.53 million year old early adolescent preserves major portions of the axial and appendicular skeletons (primarily lacking most of the hand and foot bones) (296, 297). Although developmentally similar to a modern human ≈12 year old, dental structure and development indicates an age-at-death ≈8 years (Dean). Despite being immature, the completeness of KNM- WT 15000 relative to all other pre-Late Pleistocene Homo associated remains has resulted in its being employed extensively to assess body size, proportions and other aspects of early Homo paleobiology. However, there have been suggestions that the individual suffered developmental dysplasias of the vertebral column in particular and is therefore inappropriate as a reflection of early Homo paleobiology (see discussion in 298). However, Figure S37. Inferoposterior view extensive review of its vertebral column (C7, 10 , of the KNM-WT 15000 (Nariokotome) anterior foramen L1 to S5) (298) indicates that the “unusual” aspects of its vertebrae with the occipital condyles and the are due to a combination of fossilization damage, immaturity and basioccipital. The condylus tertius ancestral morphology. The only pathological aspects are L4 and L5 is circled. Reprinted with articular facet subluxations, most likely post-traumatic (299). permission from ref. 298. At the same time, KNM-WT 15000 exhibits a minor variant of the craniocervical articulation, a supplementary articular facet on the anterior foramen magnum (or posterior basioccipital), a condylus tertius (Fig. S37). It is a normal variant of the occipito-atlanto-odontoid articulation and is manifested as a discrete facet on the basioccipital (3002). It does not reflect skeletal dysplasia but is the product of variation in embryonic development (298). It is usually asymptomatic, encountered incidentally clinically, and occasionally employed as a discrete skeletal variant. Its incidence in a pooled recent human sample (301-303) is 0.56% (n = 2,130). Its presence in the Nariokotome early adolescent is therefore unusual, given the <10 early Pleistocene crania preserving the basioccipital. Incidence: <1.0%

Rochereil 1 – Vertebral Column The Grotte de Rochereil in southwestern France yielded Late Upper Paleolithic (Magdalenian and Azilian, MIS 2) levels, with human remains deriving from both horizons (149; see Rochereil 3 above). From the Azilian level the largely complete skeleton of an older adult male (Rochereil 1) was discovered in a tightly flexed position, leading to the preservation of the skull, the major long bones and the complete vertebral column (151, 304). In particular, it has been possible to articulate the vertebrae close to their in vivo positions by carefully matching the adjacent articular facets. When so articulated, the vertebrae exhibit marked , with a left concavity from T2 to L1 and an axial rotation of the lumbar vertebrae

38 such that the ventral L5 body is oriented left relative to the thoracic column (305, 306; Fig. S38). It would be difficult to straighten the vertebral column without separating adjacent vertebral articular facets. This scoliosis is accompanied by vertebral degenerations in the cervical region and from T11 to L5. The only other lesion on the skeleton is a degeneration of the distal right radioulnar articulation, most likely post-traumatic (298). The vertebral scoliosis was Figure S38. Dorsal view of the Rochereil 1 vertebral originally (305, 307) attributed t o column, from C1 to L5. The adjacent vertebrae were a developmental asymmetry of the articulated according to close matching of their shoulders, a congenital elevation articular facets. Reprinted with permission from ref. of the shoulder (Sprendel’s 305. deformity). There is modest asymmetry in clavicular length (5.6%) and scapular height (4.1%), with the right side smaller in both cases. There is also a dorsoinferior expansion of the clavicular facets on the manubrium, suggesting a more open angle for the . However, at least the clavicular asymmetry is well within recent human normal variation (305, 308); the scapulae are narrow and tall with straight vertebral borders (304), rather than the shorter and wider scapulae with strongly convex medial sides associated with Sprengel’s deformity (309). It is therefore difficult to support a diagnosis of Sprengel’s deformity for Rochereil 1. Yet, the marked scoliosis is apparent, and it is likely to have been primarily developmental rather than degenerative given the orientations of the vertebral articular facets. The ultimate etiology of the scoliosis remains unclear. Unknown etiology

Shanidar 1 and 4 – Sacra The Middle Paleolithic levels of , in Iraqi Kurdistan, have yielded the remains 10 individuals, three of them (Shanidar 1, 3 and 5) from late Middle Paleolithic levels (MIS 3) and the remainder from earlier (MIS 4) levels (3, 310, 311). Three of the associated skeletons, Shanidar 1, 3 and 4, have yielded variably complete sacra. These sacra vary in the degrees to which their sacral canal hiatuses extend cranially from the caudal sacral apex, a trait that appears to be under substantial genetic control (312). The Shanidar 3 sacrum has its hiatus apex at S3/S4. In a pooled sample of recent human sacra (n = 1,919, and not including sacra with full ), 16.6% have it at S3 and 59.1% at S4 (313- 317), and other Neandertals exhibit a similar sacral hiatus apex level (294, 318). Shanidar 1 and 4, however, have more cranial apex positions, being at Figure S39. The Shanidar 1 (A) and 4 (B) sacra in S2 and S2/S3 respectively (Fig. S39). In a dorsal view, showing the enlarged sacral hiatus to S2 and S2/S3 respectively. Approximately to scale. geographically diverse recent human sample (n = 454) (313), 4.2% have it at S2/S3 or above, and in the larger pooled sample (n = 1,919), only 1.4% have it cranial of S2/S3. The largely open sacral hiatus of Shanidar 1 is therefore unusual among recent humans, and the slightly smaller one of Shanidar 4 is not much more common. Moreover, given the stratigraphic and chronological separation of Shanidar 1 and 4, these are independent examples of this rare sacral developmental variant. Incidences: <5.0% each

39

Shanidar 3 and Kebara 2 – L1s The superior Middle Paleolithic levels (MIS 3) of Shanidar Cave, in Iraqi Kurdistan, yielded three partial skeletons (Shanidar 1, 3 and 5), of which the second was modestly deeper in the stratigraphy than the other two (3, 319). The Shanidar 3 partial skeleton of an older male preserves, along with four teeth and fragmentary pelvic, manual, and pedal remains, a major portion of the costal skeleton, portions of T1 to T10 vertebrae, and the largely complete (with some restoration) T11 to S5 vertebral column (3, 320, 321). The T11 to L5 vertebrae are securely assigned to thoracic versus lumbar vertebrae based on their spinous process and especially articular facet morphology.

Figure S40. The Shanidar 3 T11 to L5 articulated vertebrae in left lateral view (A), and the Shanidar 3 L1 (cast) in left lateral (B) and right lateral (C) views. The arrows indicate the articular facets for the first lumbar ribs. Portions of the vertebrae, but not the relevant aspects of the L1, were restored in plaster by T.D. Stewart. Left image courtesy of Charles E. Hilton (University of North Carolina at Chapel Hill, Chapel Hill, NC). As documented by Ogilvie and colleagues (320), the Shanidar 3 L1 (vertebra 20) lacks transverse processes and exhibits distinct articular facets on the pedicles adjacent to the zygapophyses (Fig. S40). The facets appear as normal articular surfaces, and there is no evidence trauma to the L1. The other abnormalities on the Shanidar 3 vertebrae are osteoarthritic growths on thoracic laminae, T12 to L5 bodies, and L2 to L4 articular facets (3, 320). The first two lumbar vertebrae and the T12 also exhibit ventral wedging of the vertebral bodies (see below). The usual transverse processes (or preserved portions thereof) are present on the L2 to L5 vertebrae. This variant of the L1 is also present on the 20th vertebra of the complete vertebral column of the late Middle Paleolithic (MIS 3) late archaic human from Kebara Cave, northern Israel, Kebara 2 (4, 290, 293; Fig. S41). The Kebara 2 L1 has distinct rounded articular facets on the lateral surfaces of dorsal pedicles, flat on the right and convex on the left. The L1 ribs are preserved, flattened dorsoventrally, extending 19.6 and 21.2 mm laterally. The Kebara 2 vertebral column does present developmental anomalies of the more caudal lumbar vertebrae and the sacrum (see above), and it is has been Figure S41. The Kebara 2 L1 in suggested (4) that the presence of lumbar ribs on its L1 is related to cranial view (left), the separate right lumbar rib (middle), and those anomalies. The well-known, if infrequent, presence of L1 ribs right lateral view of the dorsal among recent humans suggests that it is more likely a normal variant, body and neural arch with the distinct from the unusual more caudal anomalies. lumbar rib facet indicated (right). The Kebara 2 and Shanidar 3 L1s therefore possessed lumbar Images courtesy of T. D. Berger ribs as opposed to transverse processes. None of the other sufficiently (photographer), Charles E. Hilton (University of North intact late archaic L1s (the MIS 4 Regourdou 1 and Shanidar 2 and 4) Carolina at Chapel Hill, Chapel possessed lumbar ribs, as indicated by their variably preserved L1 Hill, NC), and Baruch Arensburg transverse processes (3, 322). Ribs as opposed to integral transverse (Tel Aviv University, Tel Aviv, processes on the L1 are normally considered one reflection of cranial- Israel).

40 caudal shifts of the vertebral column, in which the presence of L1 ribs would indicate a caudal shift of the vertebral column (323, 324). The Shanidar 3 zygapophyses do not indicate a caudal shift, as the L1 facets are cranially and caudally “lumbar” and the T12 facets are cranially “thoracic” and caudally “lumbar.” The same configuration is present on the Kebara 2 T12 and L1 vertebrae. Frequencies of L1 ribs (whether unilateral or bilateral) vary from 0.5% to 1.2% in recent human samples (324-328) [Lanier (323) reported 8.8% lumbar ribs, but it is unclear whether he limited them to vertebra 20]. The presence of L1 ribs on two of the five late archaic humans (all Neandertals) retaining sufficient portions of their first lumbar vertebrae is therefore non-normal. Incidences: <5.0% each

Shanidar 3 – T12 to L3 Bodies The Shanidar 3 partial skeleton of an older male preserves the largely complete bodies of the twelfth thoracic to fifth lumbar vertebrae (see context and preservation discussion above; Fig. S40). The vertebral bodies of the T12 to L2 present craniocaudally shorter ventral bodies than dorsal bodies, resulting in dorsoventral height (or wedging) indices of 122.6, 123.9 and 117.5. A pooled recent human sample (329, 330) provides indices of 111.9 ± 6.9 for the T12 and 110.5 ± 7.2 for the L1, and 104.7 ± 5.4 for the L2 (n = 74), placing the Shanidar 3 indices 1.55, 1.86 and 2.37 standard deviations from the means. Most other recent human samples provide similar mean indices (331-334), although a pooled Inuit sample (335) has somewhat higher mean indices. The more caudal Shanidar 3 L3 to L5 have the more usual dorsoventral proportions, with indices of 101.0, 100.3 and 81.2 respectively versus 99.6 ± 5.2, 93.8 ± 5.7 and 83.2 ± 5.7 respectively for the recent human sample. In addition to the ventral wedging of the Shanidar 3 T12 to L2, the L1 exhibits a ventral flattening of the vertebral body disk surfaces, especially cranially. These two features, variably in combination with Schmorl’s nodes (which are absent from Shanidar 3), should indicate reduced and a probable kyphosis in this vertebral segment. As such, they are skeletal reflections of spinal osteochrondrosis (Scheuermann’s disease). The condition is an abnormality of the growth cartilage endplate, mostly likely due to elevated biomechanical loads on the vertebra during growth but promoted by a genetic predisposition; its etiology nonetheless remains unclear (336-338). The deformity emerges during the juvenile and adolescent years, produces thoracolumbar kyphosis, and is most evident in vertebral body wedging. It is variably reported as occurring in <1% to ≈13% of individuals (336, 337, 339). The marked T12 to L2 ventral wedging, combined with the flattening of the L1, therefore most likely indicates the presence of spinal osteochondrosis in Shanidar 3, a developmental abnormality and not a population characteristic (contra 278). The one other Neandertal non-pathological L1, from Regourdou 1, has an unexceptional wedging index of 110.1 (322). The inferred upper lumbar kyphosis of the Kebara 2 Neandertal and the Middle Pleistocene Atapuerca-SH Pelvis 1 (278) are of uncertain significance, given the marked developmental and degenerative abnormalities of their vertebral columns (4, 278). Incidence: <5.0%

El Sidrón SD-1094 and SD-1643 – C1s The late Middle Paleolithic (MIS 3) site of El Sidrón in northwestern Spain has yielded ≥2,500 skeletal elements from ≥13 individuals (256-258; see above). Among the vertebral remains, there are three atlases (C1s), SD-1094, SD1643 and SD-1605/1595. The last of these vertebrae is developmentally normal, but the other two present defects of ossification of the anterior or posterior arch (340).

41

The SD-1094 C1 consists of the right ventrolateral anterior arch and adjacent ventral portions of the articular facets (Fig S42A). The arch is truncated at the sagittal midline, between where there would normally be bilateral synchondroses. The ventrosagittal end does not have a normal synchondrosis surface but is rounded; it therefore does not represent normal-for-age immature lack of fusion. The SD-1643 C1 is ventrally complete and fused, but dorsally it presents only the left side. It lacks the right posterior arch (or lamina) due to a post-mortem break adjacent to the right articular facets (Fig. S42B). The preserved left portion of the posterior arch terminates in a tapered and truncated end, at the location of the posterior arch synchondrosis. Its anterior synchondroses are completely fused. Differential diagnosis of these atlantal elements by Ríos and colleagues (340) indicates that each of them represents a developmental failure of synchondrosis fusion.

Figure S42. Cranial views of the El Sidrón first (C1s, atlases), not to scale. The SD-1094 right anterior arch is overlain on a modern human atlas for orientation. Note the truncations of the left posterior arch of SD-1643 and of the anterior arch of SD-1094. SD-1605/1595 is developmentally normal. Images: A. Rosas. Reprinted from ref. 340. The truncated anterior arch of SD-1094 is best diagnosed as a congenital cleft of the arch, a condition that occurs in 0.09%-0.10% of recent human samples (340). It is sometimes associated with a bipartite atlas, which cannot be assessed for SD-1094 given its incompleteness. The absence of posterior arch fusion in SD-1643 does not represent a normal-for-age non-fusion, even though the incomplete transverse process epiphysis indicates an immature individual; the posterior arch almost always fuses prior to the anterior one, and the anterior arch of SD-1643 is complete synostosed. The recent human frequencies of posterior arch non-fusion range from 0.78% to 3.84%, with a pooled sample incidence of 2.55% (n = 8,644) (data summarized in 340). The developmental arch defects of the El Sidrón first cervical vertebrae therefore represent a moderately unusual condition (SD-1643) and a rare one (SD-1094). Moreover, given the small sample of sufficiently complete Neandertal C1s (La Ferrassie 1, Kebara 2, and Krapina 98 and 100/101, plus the three from El Sidrón) (11, 293, 341), it would be exceptional to find even two with such developmental non-fusions. Incidences: <0.1% and <5.0% respectively

Taforalt 11 and 27 – Sacra The late Later Stone Age () (MIS 2) site of , in northern Morocco, has yielded a number of human burials and an even larger sample of mixed human remains (342-344). Among the mixed ossuary remains, several of the sacra, including Taforalt 11 and 27, present unusual patterns of neural arch formation (342, 345; see Figs. 31 and 32 in 342; Figs. 1 and 2 in 345). The Taforalt 11 sacrum presents a sacral hiatus with an apex at the level of the S2, a degree of opening of the sacral neural arches that is present in only 1.4% of a pooled recent human sample (n = 1,919) (313-317). At the same time, the Taforalt 27 sacrum has a caudal hiatus up to the level of the S5, but its cranial neural canal is open to the level of the caudal S2. Frequency data for the cranial dehiscence include 2.38% (n = 252), 4.57% (n = 2,866) and 7.88% (n = 2,704), for a pooled frequency of 6.01% (346-348). The latter condition is therefore only moderately unusual, but the large hiatus of Taforalt 11 is otherwise known in the Pleistocene only in the Shanidar 1 sacrum (see above). Incidence: <5.0% and <10%

42

Villabruna 1 – L5 The Late Upper Paleolithic (MIS 2) site of Villabruna, northeastern Italy, yielded the well-preserved skeleton of a young adult male (based on pelvic morphology, pubic symphysis and auricular surface metamorphosis, and proximal clavicle fusion) in an elaborate burial (349, 350). In addition to minor porotic hyperostosis, tibial periostitis and an M3 carious lesion (350, 351), the individual presents two (possibly independent) developmental abnormalities, a separation (spondylolysis) of the fifth lumbar neural arch and bilateral scapular acromial bones (see below for the latter and possible association). Figure SD43: Cranial view of the Villabruna 1 fifth The Villabruna 1 L5 (Fig. S43) is complete, and it is notable for lumbar vertebrae, with the the bilateral separation of the laminae from the zygapophyses and separate laminae and transverse processes. It also exhibits an accessory left articular facet for the spinous process. Image sacral ala, which Vercellotti and colleagues (350), along with generally courtesy of Giuseppe depressed vertebral body disk surfaces, interpreted as the consequences of Vercellotti (Ohio State University, Columbus, high levels of vertebral column compression from burden carrying. OH). Bilateral (or unilateral) separation of the laminae from more ventral portions of the neural arch may result from serious trauma to the lumbar region, but it usually is at the interarticular portion of the neural arch and, especially for L5, commonly results in spondylolisthesis with consequent postural and neurological deficits (284, 352, 353). Despite the noted changes in the vertebral bodies, and given the accessory facet between the Villabruna 1 left L5 transverse process and the S1 ala, there does not appear to have been more than trivial amounts of spondylolisthesis associated with the L5 spondylolysis. The laminar separation is therefore not likely to have resulted from localized trauma. However, there are persistent associations of spondylolysis with a variety of athletic activities that stress the lumbar region (282, 285), and the incidence of spondylolysis generally increases with age (281, 285). At the same time, there are dysplastic/familial tendencies to develop spondylolysis (284, 285), suggesting predispositions for neural arch separations. The bilateral laminar spondylolysis of Villabruna 1 is therefore best considered (as noted by 350) as a developmental abnormality associated with elevated level stress on the lumbar column and a congenital predisposition for the separation. Neural arch defects of the lumbar vertebrae have been documented skeletally in 4.5% (n = 7,063) of a pooled recent human adult sample (279-282), although substantially higher frequencies have been noted in prehistoric Native American samples (283). In the largest individual recent human sample (n = 4,200), 85.2% of the separations (n = 183) were in the L5, and of those, 82.7% (n = 156) involved bilateral separations (or 3.1% bilateral L5 occurrence in the sample) (280). The presence of spondylolysis in Villabruna 1 is therefore modestly unusual for a robust Upper Paleolithic forager. Incidence: <5.0%

Shoulder and Arm Variants Barma Grande 2 – Humeri The large cave of Barma Grande, in the Balzi Rossi cliff in northwestern Italy, yielded an Upper Paleolithic sequence with four human burials, three single ones (Barma Grande 1, 5 and 6) and a triple one (Barma Grande 2 to 4) (287, 354-357). The burials lack secure stratigraphic contexts, given their “excavation” in the late nineteenth century, but a combination of the associated burial goods and a direct

43 date on Barma Grande 6 places them within the Mid Upper Paleolithic (Gravettian, MIS 3) (356, 358). The triple burial consists of an older adult male (Barma Grande 2), a younger adolescent (Barma Grande 3) and middle adolescent (Barma Grande 4). Of concern is Barma Grande 2, who is notable for his large size (even for a Gravettian male) and especially the marked asymmetry of his upper limbs (357, 359). The upper limb asymmetry of Barma Grande 2 is most evident in his humeral diaphyses (Fig. S44), in which both of them have relatively pronounced muscle markings but the right humeral diaphysis is strikingly larger than the left one. The mid-distal (35%) polar moment of area asymmetry value of 170.5 is 3.74

Figure S44. The Barma Grande 2 humeri in anterior view with the reconstructed mid-proximal (65%) diaphyseal cross- sections (left), and a comparison of the Barma Grande 2 mid-distal (35%) polar moment of area asymmetry to values from Upper Paleolithic paired humeri (right). Humeral images courtesy of Vitale S. Sparacello (Université de Bordeaux, Pessac, France) (copyright MIBAC-SABAP Liguria). Cross-sections adapted from ref. 346. The asymmetry of Barma Grande 2 (170.5%) is compared to a pan-Old World Early/Mid Upper Paleolithic sample (E/MUP) and to a Late Upper Paleolithic sample from western Eurasia; data from Sparacello et al. (51). standard deviations from an Early/Mid Upper Paleolithic mean (48.3 ± 32.7, n = 16; max: 108.1) and 3.97 standard deviations from a pooled western Eurasian Upper Paleolithic sample (48.9 ± 30.6, n = 36) (Fig. S44) (p = 0.0025 and 0.0002 respectively). The diaphyses of the radii and ulnae also exhibit consistently greater right side diameters (359). The epiphyses of the upper limb exhibit apparently lower asymmetry, especially as reflected in his humeral distal bi- epicondylar asymmetry. As a result, in their detailed assessment of the upper limb asymmetry and related paleopathology of Barma Grande 2, Churchill and Formicola (359) concluded that he had sustained a soft tissue injury or a neuropathy to the left upper limb as an adult, leading to the much weaker non-dominant arm. Reconsideration of the Barma Grande 2 upper limb asymmetry, however, suggests that it may well have developed prior to maturity and involved both of his . Comparisons of scaled humeral diaphyseal rigidity (Fig. S45) place the left humeral diaphysis within the Upper Paleolithic distribution, if among the more modest ones. The right humerus exceeds all of the others in relative strength. The distal Figure S45. Bivariate plots of humeral mid-distal (35%) (ln) humeral epiphyseal breadths are indeed polar moments of area versus (ln) humeral length times body trivially asymmetrical (1.4%), but the distal mass for Barma Grande 2 (BG2) and Eurasian samples of articular breadths are more markedly so Early/Mid Upper Paleolithic (E/MUP) and Late Upper (8.3%). The articular asymmetry value is Paleolithic (LUP) humeri, plotted separately for dominant substantially above an Early/Mid Upper (left) and non-dominant (right) humeri. Barma Grande 2 Paleolithic distribution (mean: 2.0%, n = data from Chuchill and Formicola (359); comparative data 12; maximum: 5.3%) and a pooled western from Sparacello et al. (51). Upper Paleolithic mean (2.2%, n = 24; maximum: 5.6%); it is approached only by

44 the pathological Dolní Věstonice 15 (7.4%). In addition, 10 of 12 carpal measurements are larger on the right side. Given that diaphyses are more susceptible to hypertrophy during development (360) and articulations stabilize with maturity, these comparisons suggest that the asymmetry involved both right arm hypertrophy and differential loading prior to maturity. The interpretation of the Barma Grande 2 upper limb asymmetry should therefore include differential loading during development, as well as possible influences during adulthood. The ultimate etiology, as noted by Churchill and Formicola (359), remains unclear. Probability: <0.01%

Sunghir 1 – Clavicles The Sunghir 1 skeleton of an older male adult derives from the Early/Mid Upper Paleolithic (MIS 3) site of Sunghir, in northern Russia (8, 98, 99). Sunghir 1 died traumatically (361) and was then buried in a less elaborate grave than Sunghir 2 and 3 (101). Aside from the cervical wound that caused death, Sunghir 1 presents only minor degenerative lesions, including dentoalveolar deteriorations and vertebral, manual and pedal osteoarthritis. However, he presents one aspect of body proportions which places him outside of known Late Pleistocene variation and at the limits of recent human variation; he had exceptionally long clavicles (8, 362; Fig. S46).

Figure S46. The right humerus and clavicle of Sunghir 1 in anterior and cranial views (left), and a bivariate plot of (ln) clavicle maximum length versus (ln) estimated body mass for Sunghir 1 (Su) and Late Pleistocene samples (right) (see 363). For the comparative samples, Late archaic: Neandertals; MPMH: Middle Paleolithic modern humans; E/MUP: Early/Mid Upper Paleolithic; LUP: Late Upper Paleolithic. Although his stature and body mass were among the larger of the Mid Upper Paleolithic ones known (8), the maximum length of his right clavicle (≈194 mm) is 3.66 standard deviations from a Mid Upper Paleolithic mean (154.9 ± 10.7 mm, n = 17) and 14% longer than the next longest clavicle (Dolní Věstonice 16 at 170 mm). Its claviculohumeral index of ≈53.6 is 2.44 standard deviations from the Mid Upper Paleolithic mean (46.4 ± 3.0, n = 17) and is most closely approached by those of Dolní Věstonice 16 and Mittlere Klause 1 (both 51.2) among Mid Upper Paleolithic humans. It is among the high values of ≈53.2, 53.1 and 53.3 for the Feldhofer 1, La Ferrassie 1 and Palomas 96 Neandertals. It is matched or exceeded by only 0.6% of a global recent human sample (n = 1,315) (363). However, clavicle length is best scaled relative to body mass, and there is a consistent relationship between clavicle length and body mass across Pleistocene and recent humans (363). When the Sunghir 1 clavicle length is compared to his estimated body mass (Fig. S46), it falls well above the pooled Late Pleistocene sample distribution, 2.70 standard deviations from the Late Pleistocene average. Moreover, his clavicle to body mass proportions are matched by only 0.8% of a global recent human sample (n = 1,255) (363). The Sunghir 1 clavicles, while otherwise normal, therefore have exceptional lengths, however they are scaled for body size. Incidence: <1.0%

45

Villabruna 1 – Scapulae The Late Upper Paleolithic (MIS 2) site of Villabruna, northeastern Italy, yielded the well- preserved skeleton of a young adult (mid-third decade) male (see above) (350). In addition to minor porotic hyperostosis, tibial periostitis and an M3 carious lesion (350), the individual presents two (apparently independent) developmental abnormalities, spondylolysis of the L5 neural arch (see above) and bilateral scapular acromial bones (os acromiale; Fig. S47). The bilaterally symmetrical acromial bones are the product of non- fusion of the distal acromial epiphysis, which normally fuses to the remainder of the scapula by the end of the second decade. Given the age-at-death of ≈25 years for Villabruna 1, his acromial epiphyses should have been fused; they are clearly separate (Fig. S47). The persistence of an os acromiale is frequently associated with shoulder pain and rotator cuff difficulties, but it may be asymptomatic (364, 365). In pooled recent human adult samples, both skeletal Figure S47. Cranial and radiographic, it has an incidence of 5.81% (n = 2,408) by individual, views of the Villabruna 3.76% (n = 5,857) by scapula; and 2.23% (n = 2,198) bilaterally by individual 1 scapulae, with their (364, 366-368). It therefore represents an unusual but not especially rare bilateral acromial developmental variant. Moreover, the presence of os acromiale may represent bones. Images courtesy delayed fusion as much as non-fusion. of Giuseppe Vercellotti (Ohio State University, Although probably independent, it is possible that the Villabruna 1 os Columbus, OH). acromiale and L5 spondylolysis (see above) are related, because they both reflect incomplete developmental fusion. They occasionally co-occur (0.9% in a pooled recent human sample, n = 134) (369). They are nonetheless considered separately here. Incidence: <5.0%

Pelvis and Leg Abnormalities Arene Candide 2 – Femora The Late Upper Paleolithic (final Epigravettian, MIS 2) virtually complete Arene Candide 2 skeleton of a young adult male, from the site of Arene Candide in northwestern Italy (32, 44, 45), was buried with the remains of Arene Candide 3 and 4 arranged around him (32, 276; see above). His primary long bones have been described as small (276), although his humeral maximum (302.5 and 307.5 mm) and femoral bicondylar (421 and 425 mm) lengths fall in the middles of the Late Upper Paleolithic distributions (Fig. S2), whereas his femoral head diameters (48.0 and 50.5 mm) are among the highest Late Upper Paleolithic values (Fig. Figure S48. Posterior S2). views of the right and left proximal femora of Arene At the same time, his femora Candide 2, and lateral view are exceptional in lacking lesser of the proximal right femur trochanters (276) (Fig. S48). In the (middle). Note the bilateral place of the lesser trochanters, the absence of the lesser posteromedial proximal diaphyses trochanters. Images courtesy of Vitale S. have irregular surfaces, with low Sparacello (Université de ridges of bone in the areas of the Bordeaux, Pessac, France) lesser trochanteric metaphyses and (copyright MIBAC- extending distally of those areas. SABAP Liguria). There is no evidence of the

46 metaphyses. Other than a roughening and ridges in the areas of the lesser trochanters, there is no indication of the insertions of the ilio-psoas tendons. Yet, the vertebral and internal iliac areas for the origins of psoas major and iliacus appear normal, and there are well-developed ilio-psoas sulci adjacent to the anterior inferior iliac spines. The ilio-psoas muscles therefore appear to have been present and functioning, and they therefore must have had diffuse proximal diaphyseal insertions rather than trochanteric ones. Moreover, each femur has a well-developed greater trochanter, linea aspera and intertrochanteric line, and there is no indication of hypotrophy in the femoral or tibial diaphyses (45, 276). Formicola and colleagues (276) assessed ilio-psoas paralysis, rickets, avascular necrosis, congenital trochanteric absence, and traumatic avulsion as diagnoses for the lesser trochanter absence of Arene Candide 2. An additional possibility is atraumatic avulsion (370). The first three are readily contradicted by the condition of the Arene Candide 2 remains. The fourth condition is known clinically only in cases of major malformation of the femur and apparently unknown in otherwise normally developed limb bones (276, 371). The fifth possibility is well-known, especially among juveniles to adolescents, from excessive loading of the hip in extreme postures with separation through the epiphyseal cartilage. It is usually unilateral and often reunites. The last diagnosis involves neoplasmic penetration of the epiphyseal cartilage, and it is excluded by a lack of other evidence for metastatic disease in the skeleton. Therefore, following Formicola et al. (276), the femora of Arene Candide 2 represent either 1) a rare bilateral lesser trochanter avulsion during development without reattachment or 2) an idiosyncratic bilateral absence or non-ossification of the lesser trochanter epiphysis. Unknown etiology

Berg Aukas 1 – Femur The paleontological site of Berg Aukas in northern (372, 373) has yielded a largely complete proximal half of the a human femur (Fig. S49A). Although the femur lacks stratigraphic context, its degree of fosilization and diaphyseal morphology aligns it with Late and especially Middle Pleistocene human femora (373, 374). It is unlikely to derive from either an australopith or an early modern human. The bone is a large human femur, with a femoral head diameter (57.6 mm) which exceeds the known Pleistocene Homo range (32.6–55.5 mm) and is 2.79 SD from its mean (45.9 ± 4.2 mm, n = 165) (375). Its maximum length, based on the estimation of the femoral midshaft from its changing contour, is ≈460 mm (374). The cross-sectional geometric parametersof its midshaft are also among the largest known for Pleistocene humans (374). Yet, if appropriately scaled to its estimated body mass and length (Fig. S49B), it falls well within the Pleistocene archaic and early modern human distributions, if beyond the ranges of their size variation. Its femoral neck angle, however, at 106°, is exceptionally low (Fig. S49C). It is below Figure S49. The Berg Aukas 1 femur (cast) in anterior and posterior views (A), comparison of its midshaft polar moment of area to length x body mass (E/M Pleist: Early/Middle Pleistocene) (B), and neck shaft angle comparisons (Archaic: Pleistocene non-modern Homo) (C). Comparative data in B are from 50, 374, 376-379. Sample sizes in C: Archaic: 18; Early modern: 53; Recent humans: 1,065; recent human samples as in 380.

47 the lowest known value for Pleistocene humans (110°: KNM-WT 15000 and Předmostí 4) and 3.26 SDs (p = 0.005) from the mean of a pooled archaic Homo sample (110°–127°, 121.2° ± 4.6°, n = 18). It is similarly below the lowest value for a geographically diverse sample of non-mechanized recent humans (110°–145°, 125.4° ± 5.6°, n = 1,065) (380) and 3.46 SDs (p = 0.0006) from its mean. Although variably correlated with different levels of mechanization, body proportions and subsistence patterns across recent human samples (53, 380, 381), human neck shaft angles are proximately the products of diffferential levels of joint reaction force at the hip during development (52, 53). They normally decrease to adult levels by adolescence, and therefore reflect principally differential loads during the first decade of life. Given that the Berg Aukas 1 femur is an isolated element, its body (and especially pelvic) proportions are unknown, even though the large femoral head relative to its femoral length suggest a non-linear body. It is therefore not possible to assess whether its femoral neck angle reflects developmental abnormalities elsewhere in its anatomy or exceptionally elevated immature (but not adult) activity levels. It nonetheless remains an outlier in both Pleistocene and recent human variation. Probability: <0.1%

Nazlet Khater 2 – Femora The Early Upper Paleolithic (MIS 3) site of Nazlet Khater IV, in northeastern (382), yielded two burials, the first of an adult (Nazlet Khater 1, possibly female) with a probably associated fetus/neonate and the second of an adult male (Nazlet Khater 2) (9, 383, 384). Based on full M3 formation but radiographic remnants of epiphyseal lines, Nazlet Khater 2 had an age-at-death in the mid-third decade; his male sex is securely based on cranial and especially pelvis morphology (9). His postcranial remains exhibit an extensive series of osteoarthritic degenerations, including osteophytes, articular degenerations with eburnation, and enthesopathies (9, 385). They involved most of the vertebral column, the upper limb and the proximal lower limb. The extent of the degenerations is unusual for a young adult, and it has been appropriately attributed to his activities at the lithic quarrying site of Nazlet Khater IV (385). In addition to these precocious degenerative lesions, Nazlet Khater 2 presents unusual upper versus lower limb proportions (Fig. S50). As documented by Crevecoeur (9:78), the humeri provide consistently greater stature estimates than the femora, such that the bivariate position of Nazlet Khater 2’s humeral versus femoral statures is outside of the 95% confidence interval of Late Pleistocene and recent humans. As is evident in Fig. S2, its humeral length is at the bottom of the Early/Mid Upper Paleolithic range, and its femoral length is substantially below the Early/Mid Upper Paleolithic range; together they put its Figure S50. The right humerus and humerofemoral proportions at the top of the Early/Mid Upper femur of Nazlet Khater 2 in anterior Paleolithic range (Fig. S3). In these proportions, Nazlet Khater 2 (left and center) and medial (right) falls close to the distinctly dysplastic Dolní Věstonice 15 (see views. The left humerus and femur above). are similar in size and morphology but less well preserved. Images The femoral head diameter of Nazlet Khater 2 is also small courtesy of Isabelle Crevecoeur (Fig. S50), falling substantially below both the earlier and later (Université de Bordeaux, Pessac, Upper Paleolithic distributions (Fig. S2). As a result of the France). diminutive sizes of both length and head diameter, its femora nonetheless provide it with a normal head-length index (Fig. S3).

48

His tibiae are too fragmentary to provide similar dimensions, but the estimated talar length and the mean trochlear dimension of Nazlet Khater 2 (≈44 and 25.9 mm) are ≈2.26 and 2.24 standard deviations from Early/Mid Upper Paleolithic means (53.7 ± 4.3 mm, n = 30; 31.5 ± 2.5, n = 24). They are below the minimum values (46 and 27.8 mm) for the small Dolní Věstonice 3 female. His entire lower limbs therefore appear to have been very small for an Early/Mid Upper Paleolithic individual, even though his humeral (and radial) dimensions are modest but not unusual. In addition, its femora exhibit high neck-shaft angles (131°, 132°). They place it slightly above the Early/Mid Upper Paleolithic range (Fig. S3), and reflect reduced proximal femoral loads during (especially) early development (52, 53). The Nazlet Khater 2 young adult skeleton, in addition to its pervasive degenerative lesions, presents femora that are unusually short with small and high neck-shaft angles. The last feature reflects reduced loads on the coxal articulations during immaturity, despite the apparently high loads on the individual as a young adult indicated by his osteoarthritis. The implications of the diminutive lower limbs are less clear, other than indicating developmental processes inhibiting longitudinal and articular growth. Unknown etiology

Regourdou 1 – Pelvis The Middle Paleolithic site of Regourdou in southwestern France (MIS 4) has yielded the partial skeleton from the burial of a (probably male) young adult, as well as at least a calcaneus of a second individual (386-390). Variably disturbed in situ and gradually enriched from the faunal collections, the reassembled Regourdou 1 partial skeleton retains a mandible, extensive vertebrae, most of the upper limbs, substantial portions of the lower limbs, and much of the pelvis. The majority of the Regourdou 1 remains indicate a robust and non-pathological individual (322, 391, 392). The pelvic remains, especially the articulating sacrum and ilia (Fig. S51), however, present a marked asymmetry (393, 394). The sacrum retains the S1, most of the S2 and a small right portion of the S3; it is developmentally normal with complete fusion of the synchondroses and the vertebral bodies. Yet, the left ala is substantially wider than the right one. Measurement from the midline S1 promontory to each arcuate line at the auricular surface provides an asymmetry value of ≈9%. Both ilia articulate well with the sacral auricular surfaces and provide even contours of the arcuate lines from the S1 to the anterior inferior iliac spine on the right and slightly dorsal of it on the left. Using the sacrum to define the coronal plane of the pelvis, the right arcuate line curves distinctly more ventrally than the left one. In combination with the alar asymmetry, the Regourdou 1 pelvis provides a markedly left- displaced pelvic aperture relative to the S1 and S2 bodies. Measurement from the mid-sagittal plane, as defined by the S1 Figure S51. Ventral (above) and vertebral disk surface and cranial articular facets, to each iliac cranial (below) views of the arcuate line provides an asymmetry value of ≈32%. Normal sacra Regourdou 1 articulated sacrum, and pelvic apertures provide substantially lower asymmetry values right and left ilia and left ischium. Portions of the right ischium and than Regourdou 1 (395-397). pubis are preserved separately. The Regourdou 1 pelvis, which also includes portions of Images courtesy of Bruno both ischia and a portion of the right superior pubic ramus, is Maureille (Université de Bordeaux, associated with two pieces of the right femur. The femoral Pessac, France). diaphysis exhibits pathological new bone growths on its

49 subperiosteal surface (387); its etiology and any relationship to the pelvic asymmetry are unknown. The more distal lower limb remains are non-pathological and, for the two elements preserving antimeres (talus and distal hallucal phalanx), symmetrical (length asymmetry values of 0.6% and 1.2% respectively). At least bi-iliac asymmetry is related locomotor asymmetry (398), and the changes in the right femoral diaphysis are likely to have affected leg movement. It is therefore not clear how the marked pelvic asymmetry, femoral abnormalities, and distal lower limb normal and symmetrical morphology interrelate. Regardless of the etiology and consequences of this combination in Regourdou 1, its pelvic asymmetry is exceptional. Unknown etiology

Sunghir 3 – Femora The Sunghir 3 skeleton derives from the Early/Mid Upper Paleolithic (MIS 3) site of Sunghir, in northern Russia (98, 99). Sunghir 3 is a late juvenile (≈10 years) male, based on his dentition, pelvis and aDNA (8, 100). He was buried with Sunghir 2 in the richest known Upper Paleolithic burial (8, 99, 101). Most of the very complete Sunghir 3 skeleton appears normal. The tibial diaphyses indicate normal levels of locomotor anatomy hypertrophy for a Late Pleistocene juvenile; the humeri reflect active participation in habitual activities, as is reflected in both their rigidity and asymmetry (8, 102). However, his dentition exhibits marked dental enamel hypoplasias, and his femora are exceptional (Fig. S52). The dental enamel hypoplasias indicate elevated and especially persistent systemic stress throughout his abbreviated life (103); the unusual form of the femora has defied diagnosis (8, 399-401). The Sunghir 3 femora (Fig. S52) are most notable for their largely symmetrical, evenly curved, pronounced anterior bowing. The most projecting points on the anterior diaphyses are at midshaft, rather than distal of it as in normal human femoral curvature (402). There is a pronounced pilaster through each midshaft, probably in compensation for the increased bending moments generated by the curvature. The apparent size of the pilaster is exaggerated by the mediolateral narrowing of the mid- proximal to mid-distal diaphysis. In addition, both proximal (femoral head) and distal (condylar) metaphyses exhibit sequences of broad transverse (Harris) lines, ones which parallel the persistent dental enamel hypoplasias of the dentition. Yet, the only other transverse lines are a couple of thin ones unilaterally in the distal right tibia, arguing against systemic stress to account for the

Figure S52. The abnormal femora femoral transverse lines. of Sunghir 3, in medial (outside) The distal epiphyses have bicondylar angles of 10° and and posterior (inside) views. 11°, indicating normal developmental loading and kneeing-in of Enlarged posteroanterior the legs (403). The femoral necks exhibit moderately low neck radiographs are provided of the right femoral head and neck and of angles (121°, 115°), although they are unexceptional in a Mid the left femoral distal metaphysis Upper Paleolithic context (Fig. S3); they reflect habitual locomotor and condyles, with the sequential loads similar to those of other earlier Upper Paleolithic humans. proximal and distal transverse lines The other leg and the foot remains, from the patellae to the pedal evident Radiographs courtesy of A. phalanges, are as expected in morphology and hypertrophy for a P. Buzhilova (Lomonosov Moscow healthy and active Upper Paleolithic late juvenile. State University, Moscow, Russia) and Maria Mednikova (Institute of Sunghir 3 therefore possessed a basically normal late , Moscow, Russia). juvenile skeleton, evidence of systemic stress, and markedly abnormal femora. The bilateral symmetry of the femora implies a

50 systemic disorder. The presence of abnormal morphology only in the femora argues against systemic . The only secure diagnosis is congenital bowing of the femora (401), a descriptive rather than explanatory etiology. The Sunghir 3 femora remain unique. Unknown etiology

Tianyuan 1 – Femora The site of Tianyuandong (Zhoukoudian Locality 27, northern China) yielded major portions of a human skeleton directly dated to the Early Upper Paleolithic (MIS 3) (404, 405). The remains preserve in particular the mandible, substantial portions of the bilateral long bones, a scapula, and eight each of the hand and foot bones (10). Although disturbed, it likely derives from an intentional burial (406). The remains are of an older adult, probably male based on size. The mandible exhibits substantial ante- mortem tooth loss, the tibiae have rearrangements of the muscular attachments around the soleal lines, and there are minor periarticular osteophytes. The remainder of the skeleton is normal, except for an unusual bilateral configuration of the mid-proximal to distal femoral diaphyses (10; Fig. S53A).

Figure S53. The Tianyuan 1 femora in medial view (A), medial detail of the right distal femoral diaphysis (B) and lateromedial radiograph of the right distal femoral diaphysis (C), plus posterior view of the proximoposterior right tibia (D). The arrows on the medial detail mark the edge of the divided linea aspera.

The proximal femora, from the necks to the proximal portions of the linea aspera and pilaster, are as expected for a Late Pleistocene human, with low neck shaft angles, wide gluteal tuberosities, and moderate gluteal buttresses. The right distal femoral metaphysis is normal, and the tibial midshafts indicate normal Pleistocene levels of locomotor activity. However, beginning in the mid-proximal diaphysis and extending to the popliteal area, there are non-muscular bony crests along the posterior femoral diaphyses, which reach their maximum projection in the mid-distal diaphyses (Fig. S53A). The subperiosteal bone is normal, dense cortical bone, without any indications of porousness or new bone laid on the surface. Especially in the area of their maximum development, the crests divide the muscular lines of each linea aspera, such that separate muscle lines are along the crest medially and laterally (Fig. S54B). Radiographically (Fig. S53C) it is evident that the crests are superimposed on normal femoral diaphyses, with only minor changes in cortical thicknesses and medullary diameters through the mid-distal diaphysis. The crest is best preserved on the right femur, but the intact portions of the left femur indicate bilateral symmetry of the feature. The irregularities of the soleal line regions on the proximoposterior tibial diaphyses (Fig. S53D) may well be secondary consequences of these femoral crests. If accurate, persistent pressure on the sural muscles in hyperflexion of the from the femoral crests would have affected the muscular attachments and led to the irregular bony crests. The presence of a squatting facet on the left talus indicates frequent assumption of lower limb hyperflexion. The etiology of the diaphyseal crests on the Tianyuan 1 femora remains unclear (see discussion in 10). The absence of porous new bone or any sign of a callus, as well as the normal morphologies of the other postcrania and the proximal femora, excludes most diagnoses. The bilateral symmetry implies a

51 systemic (congenital) condition, but the limitation of the primary abnormality to the femora (excluding the tibial changes) argues against a systemic syndrome. As with the Sunghir 3 femora (see above), the femoral crests of Tianyuan 1 remain unique. Unknown etiology

Manual and Pedal Variants Baousso da Torre (Bausu da Ture) – Carpals and Tarsals The site of Baousso da Torre (Bausu da Ture) in the Balzi Rossi of northwestern Italy yielded the burials of three individuals, an adult male (Baousso da Torre 1), a younger adult male (Baousso da Torre 2), and an adolescent (Baousso da Torre 3) (286, 287, 354, 407). Although not dated radiometrically, culturally the remains are ascribed to the Mid Upper Paleolithic (Gravettian; MIS 3) (287, 355, 356, 408). The pedal and manual remains of the first partial skeleton exhibit secondary ossifications, and the second one has a similar manual growth (409). The left calcaneus of Baousso da Torre 1 exhibits a pronounced bony protuberance extending distomedially from the dorsal cuboid articular facet (Fig. S54A). It represents an ossification of the dorsal calcaneonavicular (medial band of the bifurcated) Figure S54. Pedal and manual remains of Baousso da Torre (Bausu da Ture) 1 and 2. A: medial view of the left calcaneus of Baousso da Torre 1 with the ossified calcaneonavicular coalition indicated. B: proximal view of the Baousso da Torre 1 right second metacarpal with the porous dorsal bossing indicated. C: proximal view of the Baousso da Torre 2 left second metacarpal with the dorsal boss indicated; the porous area is obscured with matrix. D: radial view of the Baousso da Torre 2 trapezoid and second metacarpal in articulation with the matching dorsal bosses Images courtesy of Sébastien Villotte (Université de Bordeaux, Pessac, France). ligament and is associated with congenital calcaneonavicular coalition (409). Pooled recent human skeletal samples provide an overall frequency of 1.4% (n = 1,374) (410), although higher frequencies have been noted radiographically and in smaller cadaver samples (409). In addition, the right second metacarpal of Baousso da Torre 1 and the left one of Baousso da Torre 2 exhibit a dorsal bony swelling, or carpal boss, of the metacarpal base (Figs. S54B and S54C) In the latter case, it is accompanied by a matching boss on the trapezoid (the trapezoid is not preserved for Baousso da Torre 1) (Fig. S54D). The proximal sides of the metacarpal bosses exhibit open porosity, most evident on Baousso da Torre 1. The incidence of carpometacarpal bossing is reported as varying from ≈1% to ≈18% depending on the sample, but its etiology remains unclear (411, 412). It may be secondary to localized hand trauma or overuse, or it may be a congenital growth abnormality; its appearance in juveniles suggests a developmental contribution in at least some of the cases (411). It is not known whether the carpometacarpal bossing or calcaneonavicular coalition of Baousso da Torre 1 and/or 2 was bilateral, given the non-preservation of the relevant antimeres. However, the co- occurrence of carpometacarpal 2 bossing for both individuals, from the same level of the site, suggests that these abnormalities are congenital and that the underlying basis for at least the metacarpal growths was shared. Moreover, the co-occurrence of the carpometacarpal and intertarsal growths in one of these individuals suggests at least a systemic predisposition to their development. Incidences: <5.0% each individual

52

Dolní Vĕstonice 16 – Trapezium and Scaphoid Bones The site of Dolní Věstonice II, in southern Moravia, Czech Republic, has yielded the isolated burial of Dolní Věstonice 16 (60, 119), an older (fifth decade) male (61-63; see above). The burial is associated with an earlier Mid Upper Paleolithic (Pavlovian) assemblage (MIS 3). The Dolní Věstonice 16 upper limbs preserve extensive manual remains, including 10 carpals (lacking only the pisiforms), 9 metacarpals, and 20 of the 28 phalanges. The manual (and other appendicular) remains are almost entirely free of articular degenerations, except for the bilateral trapezioscaphoid facets (7; Fig. S55). The articulating right and left trapezial and scaphoid facets exhibit extensive erosion of the surfaces, exposure of the trabeculae, eburnation, and altered orientation relative to the adjacent intercarpal facets. The changes represent advanced osteoarthritis of these facets. However, its distribution does not resemble normal, age related changes, as is the case with his vertebral degenerations. The altered facets are bilateral, although the humeri indicate strong right-handedness (413, 414). The remainder of the intercarpal and carpometacarpal facets are normal, including the first carpometacarpal facet, the most prone of the manual articulations (415). Moreover, in cases of radiocarpal and intercarpal arthritis, the degeneration of the trapezioscaphoid facets is normally secondary to that of the radiocarpal and capitolunate articulations (416), and those other facets are normal for Dolní Věstonice 16. There is also an absence of Figure S55. Distal views evidence for skeletal trauma in the Dolní Věstonice 16 carpal remains (the of the Dolní Věstonice 16 distal radii and ulnae are too fragmentary to assess). It is therefore likely right and left scaphoid (7) that the marked and isolated osteoarthritic degenerations of the bones, with the flattening, bilateral trapezioscaphoid facets of Dolní Věstonice 16 are the result of a porosity and eburnation of developmental abnormality, likely idiopathic joint laxity. the trapezial facets (arrows). Unknown incidence

El Sidrón, Krapina and Shanidar – Scaphoid Bones The late Middle Paleolithic site of El Sidrón in northwestern Spain (MIS 3), the earlier Middle Paleolithic site of Krapina in northern (MIS 6/5) and the Middle Paleolithic levels of Shanidar Cave in Iraqi Kurdistan (MIS Figure S56. Palmar views of the three El 4-3) have yielded a number of Sidrón scaphoid bones (SDR-064, SD-258 carpal bones, in associated and SD-679b) which exhibit os centrale skeletons (Shanidar), as portions (red arrows). The bipartite SD-96 isolated elements (Krapina) or scaphoid bone possesses an articular facet secondarily associated into (red arrow) for the missing radial portion hand skeletons (El Sidrón) (3, of the bone. Reprinted from ref. 31. 11, 31). As described by Copyright (2018), with permission from Kivell and colleagues (31), Elsevier. three of the seven scaphoid bones from El Sidrón (SDR- 064, SDR-258 and SD-679b) show an os centrale persistence, for a sample frequency of 42.9% (Fig. S56). They also noted its presence on the Shanidar 3 right and left scaphoid bones. A separate or partially

53 fused os centrale in recent humans has been noted in 0.5% and 3.1% of two samples (417), but partially fused ones represent only 1.6% of the latter sample. Other late archaic human scaphoid bones provide additional examples. One is present on the isolated Krapina 200.1 left scaphoid, and they are present on all four of Shanidar remains preserving scaphoid bones, bilaterally for Shanidar 3 and 4 and on the right for Shanidar 6 and 8 (Fig. S57). Those of Shanidar 3 and 4 are prominent, whereas the other three are more modest but distinct. The three other late archaic humans with sufficiently preserved scaphoid bones, La Ferrassie 1, Kebara 2 and Regoudou 1 (387, 418, 419), lack os centrale persistence. Assuming that the six mature scaphoids from El Sidrón represent separate individuals, eight of the 15 late archaic human individuals providing sufficiently intact scaphoid bones exhibit a partially fused os centrale, or 53.3%. It has been suggested (31) that the high frequency of this feature at El Sidrón may reflect a high level of consanguinity, something indicated by their aDNA (256) and their Figure S57. Palmar views of the Shanidar 3, mandibular canine abnormalities (259). The same 4, 6 and 8 scaphoid bones (casts), and the consideration might apply to Shanidar 4, 6 and 8, who were Krapina 200.1 scaphoid bone. The arrows point to the os centrale portions. The os either buried together or in close succession at the same centrale portion on the Shanidar 6 scaphoid location in Shanidar Cave (420), but it would not include bone is damaged, and it is small on the Shanidar 3 given its substantially higher stratigraphic Shanidar 8 bone. position (3). If the scaphoid bones from El Sidrón are pooled into two (with and without os centrale persistence) and the Shanidar 4, 6 and 8 are considered as one, the late archaic human frequency remains high at 42.9%. Therefore, although a persistent os centrale is occasionally found in recent humans, its elevated frequency among western Eurasian late archaic humans is unusual. Incidences: <5.0% each individual

El Sidrón SD-96 – Scaphoid The El Sidrón SD-96 right scaphoid bone, from the late Middle Paleolithic of northwestern Spain (MIS 3), contrasts with other El Sidrón and Neandertal scaphoid bones in lacking a tubercle and exhibiting a flat and largely square articular facet in its place (31). The morphology, which also altered the trapezium-trapezoid articular facets and their orientations, reflects a bipartite scaphoid bone with the absence of the radial portion. Bipartite scaphoid bones can result from scaphoid non-union fractures, but they also occur as a congenital anomaly (31, 421). The El Sidrón SD-96 scaphoid, given its smooth and even articular radial facet (Fig. S56), appears most likely to represent a congenital non-union of separate ossifications centers. Reported frequencies for bipartite scaphoids range from 0.13% to 0.60% (31), although as noted by Doman and Marcus (421), the frequency of congenitally bipartite ones is probably considerably less than 0.5%. Incidence: <1.0%

Acknowledgments These summaries and their implications have come together as a result of conversations with colleagues too numerous to mention individually, both paleoanthropologists and bioarcheologists, and through the increasingly common paleopathological assessments of Pleistocene human remains. A

54 number of colleagues, both collaborators and curators, have made original human remains available for personal assessment and description. More directly relevant to this synthesis, a number of colleagues have provided suggestions, information, images and/or permissions that have helped with the summaries of the individual abnormalities and anomalies. They include B. Arensburg, B.M. Auerbach, T.D. Berger, M.U. Brennen, A.P. Buzhilova, S. Condemi, I. Crevecoeur, E. Delson, H. de Lumley, H. Duday, P.F. Fabbri, R.G. Franciscus, A. Gómez-Olivencia, G. Guipert, M. Haeusler, D. Henry-Gambier, C.E. Hilton, T.W. Holliday, J.-J. Hublin, S.A. Lacy, F. Martini, B. Maureille, M.B. Mednikova, A. Rosas, S. Sázelová, V.S. Sparacello, L. Tilley, A.-M. Tillier, R.L. Tompkins, G. Vercellotti, S. Villotte, J.C. Willman, and X.J. Wu. Permission to reproduce the Arene Candide skeletal material as photographed by V.S. Sparacello was granted by the Soprintendenza Archeologia, Belle Arti e Paesaggio (SABAP) per la città metropolitana di Genova e le province di Imperia, La Spezia e Savona, MIBAC, Italy. Comments and suggestions have been provided especially by V.S. Sparacello and S. Villotte. R.G. Franciscus and A. Rosas provided helpful comments on a previous version. And much of this work has been inspired by the efforts of V. Formicola to diagnose and interpret the unusual cases in the Upper Paleolithic. To all I am immensely grateful.

Table S1. Age distribution for Pleistocene human specimens with developmental abnormalities and variants, relative to samples of Middle Pleistocene to Mid Upper Paleolithic humans. Adult remains that cannot be assigned to younger versus older adults are not included. infant child juvenile adolescent younger older 0-1 1-5 6-11 12-18 19-40 40+ Develop. Abnorm. 21 6 4 6 30 8

Mid Pleistocene 0 3 2 15 21 5 Late Archaic 15 35 34 42 92 22 MPMH2 2 4 6 3 11 2 E/MUP2 15 8 14 13 31 13 Comparative Total 32 50 56 73 155 42

1 The Krems-Wachtberg 1 and 2 neonate probable twins are counted as two individuals, although they represent one abnormality. 2 MPMH: Middle Paleolithic modern humans; E/MUP: Early/Middle Upper Paleolithic modern humans.

Supporting Information - References

1. Bonnet R (1919) Die Skelete. Der Diluviale Menschenfund von Obercassel bei Bonn, eds Verworn M, Bonnet R, Steinmann G (J.F. Bergmann, Wiesbaden) pp 11-185. 2. de Lumley MA (1973) Anténéandertaliens et Néandertaliens du bassin méditerranéen occidental européen. Étud Quatern 2:1-626. 3. Trinkaus E (1983) The Shanidar Neandertals. (Academic Press, New York).

55

4. Duday H, Arensburg B (1991) La pathologie. Le Squelette Moustérien de Kébara 2, eds Bar Yosef O, Vandermeersch B (CNRS, Paris), pp 179-193. 5. Tillier AM (1999) Les Enfants Moustériens de Qafzeh. Interprétation Phylogénétique et Paléoauxologique. (CNRS, Paris). 6. Crubézy E, Henry-Gambier D (2001) Pathologie et santé. Sépulture des Enfants de Grimaldi (Baoussé- Roussé, Italie), ed Henry-Gambier D (Éditions du Comité des travaux historiques et scientifiques. Réunion des Musées Nationaux, Paris), pp 103-212. 7. Trinkaus E, Hillson SW, Franciscus RG, Holliday TW (2006) Skeletal and dental paleopathology. Early Modern Human Evolution in Central Europe: The People of Dolní Vĕstonice and Pavlov, eds Trinkaus E, Svoboda JA (Oxford University Press, New York), pp 419-458. 8. Trinkaus E, Buzhilova AP, Mednikova MB, Dobrovolskaya MV (2014) The People of Sunghir: Burials, Bodies and Behavior in the Earlier Upper Paleolithic. (Oxford University Press, New York). 9. Crevecoeur I (2008) Étude Anthropologique du Squelette du Paléolithique Supérieur de Nazlet Khater 2 (Égypte). (Leuven University Press, Leuven). 10. Shang H, Trinkaus E (2010) The Early Modern Human from , China. ( A&M University Press, College Station). 11. Trinkaus E (2016) The Krapina Human Postcranial Remains. Morphology, Morphometrics and Paleopathology. (FF Press, Zagreb). 12. de Lumley MA, Becam G, Colard T, Duplay J, Paquis P, Quatrehomme G (2018) Pathologie de l’homme du Lazaret. Les Restes Humains Fossiles de la Grotte du Lazaret, , Alpes-Maritimes, ed de Lumley MA (CNRS, Paris), pp 469-480. 13. Wu XJ, Schepartz LA, Liu W, Trinkaus E (2011) Antemortem trauma and survival in the Late Middle Pleistocene human cranium from Maba, south China. Proc Natl Acad Sci USA 108:19558-19562. doi.10.1073/pnas.1117113108. 14. Wu XJ, Xing S, Trinkaus E (2013) An enlarged parietal foramen in the late archaic Xujiayao 11 neurocranium from northern China, and rare anomalies in Pleistocene Homo. PLoS ONE 8(3):e59587. doi.10.1371/journal.pone.0059587. 15. Brennan MU (1991) Health and disease in the Middle and Upper Paleolithic of southwestern France: a bioarcheological study. Ph.D. Thesis. New York University. 16. Villotte S (2009) Enthésopathies et activités des hommes préhistoriques - Recherche méthodologique et application aux fossiles européens du Paléolithique supérieur et du Mésolithique. Brit Archaeol Rep Int Series 1992:1-206. 17. Lacy SA (2014) Oral Health and its Implications in Late Pleistocene Western Eurasian Humans. Ph.D. Thesis, Washington University. 18. Martín-Francés L (2015) Revisión y Estudio des las Manifestaciones Paleopatológicas en los Homininos del Plio-Pleistoceno, con Especial Referencia a Algunos Fósiles de la Sierra de Atapuerca. PhD Thesis, Universidad de Alcalá. 19. Ogilvie MD, Curran BK, Trinkaus E (1989) Incidence and patterning of dental enamel hypoplasia among the Neandertals. Am J Phys Anthropol 79:25-41. 20. Berger TD, Trinkaus E (1995) Patterns of trauma among the Neandertals. J Archaeol Sci 22:841-852. 21. Guatelli-Steinberg D, Larsen CS, Hutchinson DL (2004) Prevalence and the duration of linear enamel hypoplasia: a comparative study of Neandertals and Inuit foragers. J Hum Evol 47:65-84. 22. Spoor F, Stringer CB, Zonneveld F (1998) Rare temporal bone pathology of the Singa from Sudan. Am J Phys Anthropol 107:41-50. 23. Lebel S, Trinkaus E (2002) Middle Pleistocene human remains from the Bau de l’Aubesier. J Hum Evol 43:659-685. doi 10.1006/jhev.2002.0598. 24. Lordkipanidze D et al. (2005) The earliest toothless hominin skull. Nature 434:717-718. 25. Gracia A et al. (2009) in a Middle Pleistocene human: Cranium 14 from the Sima de los Huesos, Atapuerca, Spain. Proc Natl Acad Sci USA 106:6573–6578. doi.10.1073/pnas.0900965106. 26. Tilley L (2015) Accommodating difference in the prehistoric past: Revisiting the case of Romito 2 from a bioarchaeology of care perspective. Int J Paleopathol 8:64-74. doi.10.1016/j.ijpp.2014.10.003. 27. Trinkaus E, Villotte S (2017) External auditory exostoses and in the Shanidar 1 Neandertal. PLoS ONE 12(10):e0186684. doi.10.1371/journal.pone.0186684. 28. Villotte S, Ogden AR, Trinkaus E (2018) Dental abnormalities and oral pathology of the Pataud 1 Upper Paleolithic human. Bull Mém Soc Anthropol Paris. doi.10.3166/bmsap-2018-0020.

56

29. Formicola V (2007) From the Sunghir children to the Romito dwarf. Aspects of the Upper Paleolithic funerary landscape. Curr Anthropol 48:446-453. 30. Formicola V (1995) X-linked hypophosphatemic rickets: a probable Upper Paleolithic case. Am J Phys Anthropol 98:403-409. 31. Kivell TL et al. (2018) New Neandertal wrist bones from El Sidrón, Spain (1994-2009). J Hum Evol 114:45-75. doi.10.1016/j.jhevol.2017.09.007. 32. Sparacello VS, Rossi S, Pettitt P, Roberts CA, Riel-Salvatore J, Formicola V (2018) New insights on Final Epigravettian funerary behavior at Arene Candide Cave (Western Liguria, Italy). J Anthropol Sci 96:1-24. 33. Bräuer G (1988) Osteometrie. Anthropologie I. ed Knussman R (Fischer Verlag, Stuttgart) pp160-232. 34. Howells WW (1973) Cranial Variation in Man. Pap Peabody Mus 67:1-259. 35. Auerbach BM, Ruff CB (2006) Limb bone bilateral asymmetry: variability and commonality among modern humans. J Hum Evol 50:203-218. doi:10.1016/j.jhevol.2005.09.004. 36. Wood B, Collard M (1999) The human genus. Science 284:65-71. 37. Brown P et al. (2004) A new small-bodied hominin from the Late Pleistocene of Flores, Indonesia. Nature 431:1055-1061. 38. Morwood MJ et al. (2005) Further evidence for small-bodied hominins from the Late Pleistocene of Flores, Indonesia. Nature 437:1012-1017. 39. van den Bergh GD et al. (2016) Homo floresiensis-like fossils from the early Middle Pleistocene of Flores. Nature 534:245-249. 40. Berger LR et al. (2015) Homo naledi, a new species of the genus Homo from the Dinaledi Chamber, . eLife 4:e09560. doi.10.7554/eLife.09560. 41. Hawks J et al. (2017) New fossil remains of Homo naledi from the Lesedi Chamber, South Africa. eLife 6:e24232. doi.10.7554/eLife.24232. 42. Jungers WL et al. (2009) The foot of Homo floresiensis. Nature 459:81-84. 43. Kivell TL et al. (2015) The hand of Homo naledi. Nature Comm 6:8431. doi.10.1038/ncomms9431. 44. Cardini L (1980) La necropoli mesolitica delle Arene Candide. Mem Ist Ital Paleontol Umana 3:9-31. 45. Paoli G, Parenti R, Sergi S (1980) Gli scheletri mesolitici della Caverna delle Arene Candide (Liguria). Mem Ist Ital Paleontol Umana 3:33-154. 46. Sládek V, Trinkaus E, Hillson SW, Holliday TW (2000) The People of the Pavlovian: Skeletal Catalogue and Osteometrics of the Gravettian Fossil Hominids from Dolní Vĕstonice and Pavlov. Dolní Vĕstonice Studies 5:1-244. 47. Villotte S, Santos F, Courtaud P (2015) In situ study of the Gravettian individual from Cussac Cave, Locus 2 (Dordogne, France). Am J Phys Anthropol 158:759-768. 48. Feldesman MR, Kleckner JG, Lundy JK (1990) Femur/stature ratio and estimates of stature in mid- and late-Pleistocene fossil hominids. Am J Phys Anthropol 83:359-372. 49. Ruff CB, Burgess ML, Squyres N, Junno JA, Trinkaus E (2018) Lower limb articular scaling and body mass estimation in and Pleistocene humans. J Hum Evol 115:85-111. doi.10.1016/j.jhevol.2017.10.014. 50. Holt BM (1999) Mobility in Upper Paleolithic and Europe: biomechanical evidence from the lower limb. Ph.D. Dissertation, University of Missouri. 51. Sparacello VS, Villotte S, Shackelford LL, Trinkaus E (2017) Patterns of humeral asymmetry among Late Pleistocene humans. C R Palevol 16:680-689. doi.10.1016/j.crpv.2016.09.001. 52. Anderson JY, Trinkaus E (1998) Patterns of sexual, bilateral and inter-populational variation in human femoral neck-shaft angles. J Anat 192:279-285. 53. Child SL, Cowgill LW (2017) Femoral neck-shaft angle and climate-induced body proportions. Am J Phys Anthropol 164:720-735. doi.10.1002/ajpa.23311. 54. Carpenter TO, Imel EA, Holm IA, Jan de Beur SM, Insogna KL (2011) A clinician’s guide to X-linked hypophosphatemia. J Bone Min Res 26(7):1381-1388. doi.10.1002/jbmr.340. 55. Scriver CR, Beaudet AL, Sly WS, Valle D (1995) MMBID-7 summary table. The Metabolic and Molecular Bases of Inherited Disease, 7th ed, eds Scriver CR, Beaudet AL, Sly WS, Valle D (McGraw-Hill, New York), pp 1-50. 56. Aujoulat N et al. (2001) La grotte ornée de Cussac (Dordogne). Observations liminaires. Paléo 13:9-18. 57. Henry-Gambier D et al. (2013) Grotte de Cussac (Le Buisson-de-Cadouin, Dordogne): un example de comportement original pour le Gravettien. Congr Préhist France 27:169-182. 58. Klíma B (1987) Une triple sépulture du Pavlovien à Dolní Vĕstonice, Tchécoslovaquie. L’Anthropol 91:329-334.

57

59. Trinkaus E, Svoboda JA eds (2006) Early Modern Human Evolution in Central Europe: The People of Dolní Vĕstonice and Pavlov. (Oxford University Press, New York). 60. Svoboda JA (2016) Dolní Věstonice – Pavlov. (Academia, Prague). 61. Hillson SW, Franciscus RG, Holliday TW, Trinkaus E (2006) The ages-at-death. Early Modern Human Evolution in Central Europe: The People of Dolní Vĕstonice and Pavlov, eds Trinkaus E, Svoboda JA (Oxford University Press, New York), pp 31-45. 62. Brůžek J, Franciscus RG, Novotný V, Trinkaus E (2006) The assessment of sex. Early Modern Human Evolution in Central Europe: The People of Dolní Vĕstonice and Pavlov, eds Trinkaus E, Svoboda JA (Oxford University Press, New York), pp 46-62. 63. Mittnik A, Krause J (2016) Genetic analysis of the Dolní Věstonice human remains. Dolní Věstonice II : Chronostratigraphy, Paleoethnology, Paleoanthropology, ed Svoboda JA. Dolní Věstonice Studies 21:377- 384. 64. Trinkaus E (2011) The postcranial dimensions of the La Chapelle-aux-Saints 1 Neandertal. Am J Phys Anthropol 145:461-468. doi.10.1002/ajpa.21528. 65. Trinkaus E (2006) The lower limb remains. Early Modern Human Evolution in Central Europe: The People of Dolní Vĕstonice and Pavlov, eds Trinkaus E, Svoboda JA (Oxford University Press, New York), pp 380-418. 66. Trinkaus E, Formicola V, Svoboda J, Hillson SW, Holliday TW (2001) Dolní Vĕstonice 15: Pathology and persistence in the Pavlovian. J Archaeol Sci 28:1291-1308. doi.10.1006/jasc.2001.0678. 67. Hillson SW (2006) Dental morphology, proportions and attrition. Early Modern Human Evolution in Central Europe: The People of Dolní Vĕstonice and Pavlov, eds Trinkaus E, Svoboda JA (Oxford University Press, New York), pp 179-223. 68. Lovejoy CO, Heiple KG (1981) The analysis of fractures in skeletal populations with an example from the Libben Site, Ottawa County, Ohio. Am J Phys Anthropol 55:529-541. 69. Jurmain R (2001) Paleoepidemiological patterns of trauma in a prehistoric population from central California. Am J Phys Anthropol 115:13-23. 70. Cooper C, Dennison EM, Leufkens HGM, Bishop N, van Staa TP (2004) Epidemiology of childhood fractures in Britain: A study using the general practice research database. J Bone Min Res 19:1976-1981. doi.org/10.1359/jbmr.040902 71. Ludewig PM, Borstad JD (2011) The shoulder complex. Joint Structure and Function, 5th ed, eds Levangie PK, Norkin CC (F.A. Davis, Philadelphia), pp 231-270. 72. Vlček E (1991) Die Mammutjäger von Dolní Vĕstonice. Archäologie Museum 22:1-136. 73. Jelínek J (1992) New Upper Palaeolithic burials from Dolni Věstonice. Étud Rech Archéol Univ Liège 56:207-228. 74. Formicola V, Pontrandolfi A, Svoboda J (2001) The Upper Paleolithic triple burial of Dolní Vĕstonice: Pathology and funerary behavior. Am J Phys Anthropol 115:372-379. 75. Einwögerer T, Friesinger H, Händel M, Neugebauer-Maresch C, Simon U, Teschler-Nicola M (2006) Upper Palaeolithic infant burials. Nature 444:285. 76. Heim JL (1982) Les Enfants Néandertaliens de La Ferrassie. (Masson, Paris). 77. Hall JG (2003) Twinning. Lancet 362:735-743. 78. Geissmann T (1990) Twinning in catarrhine primates. Hum Evol 5(4):387-396. 79. Vandermeersch B (1981) Les Hommes Fossiles de Qafzeh (Israël). (CNRS, Paris). 80. Valladas H, Reyes JL, Joron JL, Valladas G, Bar-Yosef O, Vandermeersch B (1988) Thermoluminescence dating of “Proto-Cro-Magnon” remains from Israel and the origin of modern man. Nature 331:614-616. 81. Tillier AM, Arensburg B, Duday H, Vandermeersch B (2001) An early case of : the Middle Paleolithic Qafzeh 12 child (Israel). Am J Phys Anthropol 114:166-170. 82. Hauser G, DeStefano CF (1989) Epigenetic Variants of the Human Skull. (E. Schweizerbart’sche Verlagsbuchandlung, Stuttgart). 83. Coqueugniot H, Hublin JJ, Veillon F, Houët F, Jacob T (2004) Early brain growth in Homo erectus and implications for cognitive ability. Nature 431:299-302. 84. McCown TD, KeithA (1939) The Stone Age of Mount Carmel II: The Fossil Human Remains from the Levalloiso-Mousterian. (Clarendon Press, Oxford). 85. Richards GD, Anton SC (1991) Craniofacial configuration and postcranial development of a hydrocephalic child (ca. 2500 B.C.–500 A.D.): With a review of cases and comment on diagnostic criteria. Am J Phys Anthropol 85(2):185-200. doi.org/10.1002/ajpa.1330850207.

58

86. MacMahon B, Pugh TF, Ingalls TH (1953) Anencephalus, spina bifida, and hydrocephalus. Incidence related to sex, race, and season of birth and incidence in siblings. Brit J Prevent Soc Med 7(4):211-219. 87. Fernell E, Hagberg G, Hagberg B (1994) Infantile hydrocephalus epidemiology: an indicator of enhanced survival. Arch Dis Childhood 70:F123-F128. 88. Mallegni F, Fabbri PF (1995) The human skeletal remains from the Upper Palaeolithic burials found in Romito Cave (Papasidero, Cosenza, Italy). Bull Mém Soc Anthropol Paris 7(3-4):99-137. 89. Craig OE et al. (2010) Stable isotope analysis of Late Upper Paleolithic human and faunal remains from Grotta del Romito (Cosenza), Italy. J Archaeol Sci 37:2504-2512. doi.10.1016/j.jas.2010.05.010. 90. Fabbri PF, Lo Vetro D, Martini F (2011) Le sepolture. Grotta del Romito a Papsidero. Uomo, ambiente e culture nel Paleolitico della Calabria. Ricerche 1961–2011, eds Martini F, Lo Vetro D (Editoriale Progetto 2000, Cosenza), pp 43–53. 91. Martini F, Lo Vetro D eds (2011) Grotta del Romito a Papsidero. Uomo, ambiente e culture nel Paleolitico della Calabria. Ricerche 1961–2011. (Editoriale Progetto 2000, Cosenza). 92. Messeri P (1966) Note paleopatologiche sul materiale scheletrico umano rinvenutonella grotta del Romito a Papasidero in Calabria (Cosenza). Riun Sci Ist Ital Preist Protost 10:301–307. 93. Frayer DW, Horton WA, Macchiarelli R, Mussi M (1987) Dwarfism in an adolescent from the Italian late Upper Paleolithic. Nature 330:60-62. 94. Frayer DW, Macchiarelli R, Mussi M (1988) A case of chondrodystrophic dwarfism in the Italian Late Upper Paleolithic. Am J Phys Anthropol 75:549-565. doi.10.1002/ajpa.1330750412. 95. Fabbri PF, Graziosi P, Guerri M, Mallegni F (1989) Les hommes des sépultures de la Grotte du Romito à Papasidero (Cosenza, Italie). Hominidae, ed Giacobini G (Editoriale Jaca Book, Milan), pp 487-494. 96. Langer LG, Garrett RT (1980) Acromesomelic dysplasia. Pediatr Radiol 137:349-355. 97. Khan S, Ali RH, Abbasi S, Nawaz M, Muhammad N, Ahmad W (2012) Novel mutations in natriuretic peptide receptor-2 gene underlie acromesomelic dysplasia, type maroteaux. BMC Med Gen 13:44. doi.10.1186/1471-2350-13-44. 98. Bader ON (1978) Sungir’. An Upper Paleolithic Site (in Russian). (NAUKA, Moscow). 99. Bader ON (1998) Sungir. Palaeolithic burials (in Russian). Upper Palaeolithic Site Sungir (graves and environment), ed Bader NO (Scientific World, Moscow) pp 5-160. 100. Sikora M et al. (2017) Ancient genomes show social and reproductive behavior of early Upper Paleolithic foragers. Science 358:659-662. 10.1126/science.aao1807. 101. Trinkaus E, Buzhilova AP (2018) Diversity and differential disposal of the dead at Sunghir. Antiquity 92(361):7-21. doi.10.15184/aqy.2017.223. 102. Cowgill LW, Mednikova MB, Buzhilova AP, Trinkaus E (2015) The Sunghir 3 Upper Paleolithic juvenile: Pathology and persistence in the Paleolithic. Int J Osteoarchaeol 25:176-187. doi.10.1002/oa.2273. 103. Guatelli-Steinberg D, Buzhilova AP, Trinkaus E (2013) Developmental stress and survival among the Mid Upper Paleolithic Sunghir children: Dental enamel hypoplasias of Sunghir 2 and 3. Int J Osteoarchaeol 23:421-431. doi.10.1002/oa.1263. 104. Villotte S, Knüsel CJ (2013) Understanding entheseal changes: definition and life course changes. Int J Osteoarchaeol 23:135-146. doi.10.1002/oa.2289. 105. Formicola V, Scarsini C (1987) Contribution to the knowledge of the late Epigravettian human remains from Arene Candide cave (Liguria, Italy): a peculiar-shaped skull. Homo 38:160-170. 106. Messeri P (1980) Un cranio mesolitico ligure intenzionalmente deformato alle Arene Candide nel Finalese. Arch Antropol Etnol 109-110:497-510. 107. Anton SC (1989) Intentional cranial vault deformation and induced changes of the cranial base and face. Am J Phys Anthropol 79(2):253-267. doi.10.1002/ajpa.1330790213. 108. Pomeroy E, Stock JT, Zakrzewski SR, Lahr MM (2010) A metric study of three types of artificial cranial modification from north-central . Int J Osteoarchaeol 20:317-334. doi.10.1002/oa.1044. 109. Rhode MP, Arriaza BT (2006) Influence of cranial deformation on facial morphology among prehistoric south central Andean populations. Am J Phys Anthropol 130:462-470. doi.10.1002/ajpa.20333. 110. Delashaw JB, Persing JA, Broaddus WC, Jane JA (1989) Cranial vault growth in craniosynostosis. J Neurosurg 70:159-165. 111. Shuper A, Merlob P, Grunebaum M, Reisner SH (1985) The incidence of isolated craniosynostosis in the newborn infant. Am J Dis Child 139:85-86. 112. French LR, Jackson IT, Melton LJ (1990) A population-based study of craniosynostosis. J Clin Epidemiol 43(1):69-73.

59

113. Cornelissen M et al. (2016) Increase of prevalence of craniosynostosis. J Cran Max Facial Surg 44:1273- 1279. doi.10.1016/j.jcms.2016.07.007. 114. Hunter AGW, Rudd NL (1976) Sagittal synostosis: Its and associated clinical findings in 214 patients who lacked involvement of the coronal suture(s). 14:185-194. 115. Arsuaga JL et al. (2014) Neandertal roots: Cranial and chronological evidence from Sima de los Huesos. Science 344:1358-1363. doi.10.1126/science.1253958. 116. Arsuaga JL et al. (2015) Postcranial morphology of the Middle Pleistocene humans from Sima de los Huesos, Spain. Proc Natl Acad Sci USA 112:11524-11529. doi.10.1073/pnas.1514828112. 117. Aranburu A, Arsuaga JL, Sala N (2017) The stratigraphy of the Sima de los Huesos (Atapuerca, Spain) and implications for the origin of the fossil hominin accumulation. Quatern Int 433:5-21. doi.10.1016/j.quaint.2015.02.044. 118. Gracia A, Martínez-Lage JF, Arsuaga JL, Martínez I, Lorenzo C, Pérez-Espejo MA (2010) The earliest evidence of true lambdoid craniosynostosis: the case of “Benjamina”, a child. Child Nerv Syst 26(6):-727. doi.10.1007/s00381-010-1133-y. 119. Svoboda JA (2006) The archeological contexts of the human remains. Early Modern Human Evolution in Central Europe: The People of Dolní Vĕstonice and Pavlov, eds Trinkaus E, Svoboda JA (Oxford University Press, New York), pp 9-14. 120. Franciscus RG, Vlček E (2006) The cranial remains. Early Modern Human Evolution in Central Europe: The People of Dolní Vĕstonice and Pavlov, eds Trinkaus E, Svoboda JA (Oxford University Press, New York), pp 63-152. 121. Franciscus RG, Vlček E, Trinkaus E (2006) The mandibular remains. Early Modern Human Evolution in Central Europe: The People of Dolní Vĕstonice and Pavlov, eds Trinkaus E, Svoboda JA (Oxford University Press, New York), pp 156-178. 122. Latham RA (1971) The development, structure and growth pattern of the human mid-palatal suture. J Anat 108(1):31-41. 123. Vanderas AP (1987) Incidence of cleft lip, cleft palate, and cleft lip and palate among races: A review. Cleft Palate J 24(3):216-225. 124. Trinkaus E, Walker MJ eds (2017) The People of Palomas: The Neandertals from the Sima de las Palomas del Cabezo Gordo, Southeastern Spain. (Texas A&M University Press, College Station). 125. Walker MJ et al. (2008) Late Neandertals in southeastern Iberia: Sima de las Palomas del Cabezo Gordo, Murcia, Spain. Proc Natl Acad Sci USA 105:20631-20636. 126. Walker MJ, López MV, Haber M, Trinkaus E (2017) The context of the Sima de las Palomas Neandertals. The People of Palomas: The Neandertals from the Sima de las Palomas del Cabezo Gordo, Southeastern Spain. eds Trinkaus E, Walker MJ (Texas A&M University Press, College Station), pp 4-18. 127. Walker MJ, Lombardi AV, Zapata J, Trinkaus E (2010). Neandertal mandibles from the Sima de la Palomas del Cabezo Gordo, Murcia, southeastern Spain. Am J Phys Anthropol 142:261-272. doi.10.1002/ajpa.21223. 128. Trinkaus E et al. (2017) The Palomas mandibles. The People of Palomas: The Neandertals from the Sima de las Palomas del Cabezo Gordo, Southeastern Spain, eds Trinkaus E, Walker MJ (Texas A&M University Press, College Station), pp 30-51. 129. Rosas A (2001) Occurrence of features in mandibles from the Atapuerca-SH site. Am J Phys Anthropol 114:74-91. doi.10.1002/1096-8644(200101)114:1<74::AID-AJPA1007>3.0.CO;2-U 130. Capitan L, Peyrony D (1909) Deux squelettes humains au milieu de foyers de l’époque moustérienne. C r Acad Inscript Belles-Lettres 1909:797-806. 131. Patte E (1957) L’Enfant Néanderthalien du Pech de l’Azé. (Masson, Paris). 132. Soressi M, Jones HL, Rink WJ, Maureille B, Tillier AM (2007) The Pech-de-l’Azé I Neandertal child: ESR, uranium-series, and AMS 14C dating of its MTA type B context. J Hum Evol 52:455-466. 133. Tillier AM (1983) Le crâne d’enfant d’Engis 2: un exemple de distribution des caractères juvéniles, primitifs et néandertaliens. Bull Soc roy belge Anthropol Préhist 94:51-75. 134. Tillier AM (1996) The Pech de l’Azé and Roc de Marsal children (Middle Paleolithic, France): Skeletal evidence for variation in Neanderthal ontogeny. Hum Evol 11(2):113-119. 135. Madre-Dupouy M (1992) L’Enfant du Roc de Marsal. (C.N.R.S. Éditions, Paris). 136. Pap I, Tillier AM, Arensburg B, Chech M (1996) The Subalyuk Neanderthal remains (Hungary): a re- examination. Ann Hist-Nat Mus Nat Hung 88:233-270.

60

137. DodoY, Kondo O, Nara T (2002) The skull of the Neanderthal child of burial No. 1. Neanderthal Burials. Excavations of the Dederiyeh Cave, Afrin, Syria, eds Akazawa T, Muhesen S (International Research Center for Japanese Studies, Kyoto), pp 93-138. 138. Ishida H, Kondo O (2002) The skull of the Neanderthal child of burial No. 2. Neanderthal Burials. Excavations of the Dederiyeh Cave, Afrin, Syria, eds Akazawa T, Muhesen S (International Research Center for Japanese Studies, Kyoto), pp 271-298. 139. Vlček E (1970) Étude comparative onto-phylogénétique de l’enfant du Pech-de-l’Azé par rapport à d’autres enfants néandertaliens. Arch Inst Paléontol Hum 33:149-178. 140. Ponce de León M, Zollikofer CPE (2001) Neanderthal cranial ontogeny and its implictions for late hominid diversity. Nature 412:534-538. doi.10.1038/35087573. 141. Legoux P (1969) Détermination de l’âge dentaire de l’enfant néandertalien du Pech de l’Azé (Dordogne). C R Acad Sci Paris 268-D:2875-2878. 142. Legoux P (1970) Étude odontologique de l’enfant néandertalien du Pech-de-l’Azé. Arch Inst Paléontol Hum 33:53-87. 143. Ferembach D (1970) Le crâne de l’enfant du Pech-de-l’Azé. Arch Inst Paléontol Hum 33:13-51. 144. Ubelaker DH (1978) Human Skeletal Remains. Excavation, Analysis, Interpretation. (Aldine, Chicago). 145. AlQahtani SJ, Hector MP, Liversidge MH (2010) The London atlas of human tooth development and eruption. Am J Phys Anthropol 142:481–490. doi.10.1002/ajpa.21258. 146. Coqueugniot H (1999) Le crâne d’Homo sapiens en Eurasie: croissance et variation depuis 100 000 ans. Brit Archaeol Rep Int Ser 822:1-197. 147. Williams PL, Warwick R (1980) Gray’s Anatomy, 36th ed. (W.B. Saunders, Philadelphia). 148. Guatelli-Steinberg D (2009) Recent studies of dental development in Neandertals: Implications for Neandertal life histories. Evol Anthropol 19:9-20. doi.10.1002/evan.20190. 149. Jude PE (1960) La Grotte de Rochereil. Station Magdalénienne et Azilienne. Arch Inst Paléontol Hum 30:5-74. 150. Vallois HV (1971) Le crâne trépané magdalénien de Rochereil. Bull Soc Préhist Franç 68(2):485-495. 151. Ferembach D (1974) Le squelette humain Azilien de Rochereil (Dordogne). Bull Mém Soc Anthropol Paris 13(2):271-291. 152. Mafart B, Guipert G, Alliez-Philip C, Brau JJ (2007) Virtual reconstruction and new palaeopathological study of the Magdalenian child’s skull of Rochereil. C R Palevol 6:569-579. doi.10.1016/j.crpv.2007.09.019. 153. Mafart B (2009) Une reconstitution surprenante d’un fossile humain: la mandibule magdalénienne du crâne d’enfant Rochereil III. C R Palevol 8:403-412. doi.10.1016/j.crpv.2009.01.002. 154. Martin RD (1990) Primate Origins and Evolution. (Princeton University Press, Princeton). 155. Krogman WH (1941) Growth of Man. (Uitgeverij Dr. W. Junk, The Hague). 156. Ponce de León M et al. (2008) Neanderthal brain size at birth provides insights into the evolution of human life history. Proc Natl Acad Sci USA 105(37):13764-13768. doi.10.1073/pnas.0803917105. 157. Jaeger JJ (1981) Les hommes fossiles du Pleistocène moyen du Maghreb dans leur cadre géologique, chronologique, et paléontologique. Homo erectus. Papers in Honor of Davidson Black, eds Sigmon BA, Cybulski JS (University of Toronto Press, Toronto), pp 159-187. 158. Hublin JJ (1991) L’émergence des Homo sapiens archaïques: Afrique du Nord-Ouest et Europe Occidentale. Thèse de Doctorat d’État, Université de Bordeaux I. 159. Hublin JJ (1985) Human fossils from the North African Middle Pleistocene and the origin of Homo sapiens. Ancestors: The Hard Evidence, ed Delson E (Alan R. Liss, New York), pp 283-288. 160. Hublin JJ (2009) The of compassion. Proc Natl Acad Sci USA. 106(16):6429-6430. doi.10.1073_pnas.0902614106. 161. Cheng JCY, Tang SP, Chen TMK, Wong MWN, Wong EMC (2000) The clinical presentation and outcome of treatment of congenital muscular torticollis in infants – a study of 1,086 cases. J Pediatr Surg 35(7):1091-1096. doi.10.1053/js.2000.7833. 162. Do TT (2006) Congenital muscular torticollis: current concepts and review of treatment. Curr Opinion Pediatr 18:26-29. doi.10.1097/01.mop.0000192520.48411.fa. 163. Dunn PM (1974) Congenital sternomastoid torticollis: an intrauterine postural deformity. Arch Dis Childhood 49(10):824-825. PMC1649189. 164. Cheng JCY, Au AWY (1994) Infantile torticollis: a review of 624 cases. J Pediatr Orthop 14:802-808. PMID:7814599. 165. Douglas MT (1991) Wryneck in the ancient Hawaiians. Am J Phys Anthropol 84:261-281.

61

166. McDermott F, Stringer C, Grün R, Williams CT, Din VK, Hawkesworth CJ (1996) New Late-Pleistocene uranium-thorium and ESR dates for the Singa hominid (Sudan). J Hum Evol 31:507-516. 167. Smith Woodward A (1938) A fossil skull of an ancestral Bushman from the Anglo-Egyptian Sudan. Antiquity 12(46):190-195. doi.10.1017/S0003598X00013661. 168. Wells LH (1951) The fossil human skull from Singa. Fossil Mamm Afr 2:29-42. 169. Brothwell DR (1974) The Upper Pleistocene Singa skull: A problem in palaeontological interpretation. Bevoelkerungsbiologie, eds Bernhard W, Kandler A (Carl Fischer, Stuttgart), pp 534-545. 170. Stringer CB (1979) A re-evaluation of the fossil human calvaria from Singa, Sudan. Bull Br Mus Nat Hist (Geol) 32(1):77-83. 171. Stringer CB, Cornish L, Stuart-Macadam P (1985) Preparation and further study of the Singa skull from Sudan. Bull Br Mus Nat Hist (Geol) 38:347-358. 172. Bräuer G (2008) The origin of modern anatomy: By speciation or intraspecific evolution? Evol Anthropol 17:22-37. doi.10.1002/evan.20157. 173. Jane JA Jr, Lin KY, Jane JA Sr (2000) Sagittal synostosis. Neurosurg Focus 9(3):E3. 174. Walker PL, Bathurst RR, Richman R, Gjardrum T, Andrushko VA (2009) The causes of porotic hyperostosis and cribra orbitalia: A reappraisal of the iron-deficiency-anemia hypothesis. Am J Phys Anthropol 139:109-125. doi.10.1002/ajpa.21031. 175. Chia LP, Wei C (1976) A Palaeolithic site at Hsü-Chia-Yao in Yangkao County, Shansi Province (in Chinese with English summary). Acta Archaeol Sinica 2:97–114. 176. Norton CJ, Gao X (2008) Hominin-carnivore interactions during the Chinese Early Paleolithic: taphonomic perspectives from Xujiayao. J Hum Evol 55:164-178. doi.10.1016/j.jhevol.2008.02.006. 177. Ma N, Pei SW, Gao X (2011) A preliminary study on the stone artifacts excavated from Locality 74093 of the Xujiayao Site in 1977 (in Chinese with English summary). Acta Anthropol Sinica 30(3):275-288. 178. Wu XJ, Crevecoeur I, LiuW, Xing S, Trinkaus E (2014) Temporal labyrinths of eastern Eurasian Pleistocene humans. Proc Natl Acad Sci USA 111:10509-10513. doi.10.1073/pnas.1410735111. 179. Wei Q, Wu XJ (2012) Approach to stratigraphy of the Xujiayao-Houjiayao site in Nihewan Basin, China (in Chinese with English summary). Acta Anthropol Sinica 31(2):151-163. 180. Xing S, Martinón-Torres M, Bermúdez de Castro JM, Wu XJ, Liu W (2015) Hominin teeth from the early Late Pleistocene site of Xujiayao, northern China. Am J Phys Anthropol 156:224-240. doi.10.1002/ajpa.22641. 181. Ao H, Liu CR, Roberts AP, Zhang P, Xu XM (2017) An updated age for the Xujiayao hominin from the Nihewan Basin, North China: Implications for Middle Pleistocene human evolution in early Asia. J Hum Evol 106:54-65. doi.10.1016/j.jhevol.2017.01.014. 182. Chia LP, Wei Q, Li CR (1979) Report on the excavation of Hsuchiayao man site in 1976 (in Chinese with English summary). Vertebrata PalAsiatica 17:277-293. 183. Wu M (1986) Study of temporal bone of Xujiayao Man (in Chinese with English summary). Acta Anthropol Sinica 5:220–226. 184. Wu XJ, Trinkaus E (2014) The Xujiayao 14 mandibular ramus and Pleistocene Homo mandibular variation. C R Palevol 13:333-341. doi.10.1016/j.crpv. 185. Dharwal K (2012) Foramina parietalia permagna: the ins and outs. Folia Morphol 7:78–81. 186. Fink AM, Maixner W (2006) Enlarged parietal foramina: MR imaging features in the fetus and neonate. Am J Neuroradiol 27:1379–1381. 187. Currarino G (1976) Normal variants and congenital anomalies in the region of obelion. Am J Roentgenol 127:487–494 188. O’Rahilly R, Twohig MJ (1952). Foramina parietalia permagna. Am J Roentgenol 67:551–561. 189. Hattab FN, Yassin OM, Rawashdeh MA (1994) Supernumerary teeth: Report of three cases and review of the literature. J Dent Children 61:382-393. 190. Scheiner MA, Sampson WJ (1997) Supernumerary teeth: A review of the literature and four case reports. Australian Dent J 42(3):160-165. 191. Leco MI, Martín JF, Martínez JM (2007) An observational study of the frequency of supernumerary teeth in a population of 2000 patients. Med Oral Patol Oral Cirugia Bucal 12:e134-138. doi.10.4317/medoral.12.e134. 192. Celikoglu M, Kamak H, Oktay H (2010) Prevalence and characteristics of supernumerary teeth in a non- syndrome Turkish population: Associated pathologies and proposed treatment. Med Oral Patol Oral Cir Bucal 15(1):e575-578. doi.10.4317/medoral.15.e575.

62

193. Rajab LD, Hamdan MAM (2002) Supernumerary teeth: review of the literature and a survey of 152 cases. Int J Paediatr Dentist 12:244-254. 194. Lavelle CLB, MooreWJ (1973) The incidence of agenesis and polygenesis in the Primate dentition. Am J Phys Anthropol 38(3):671-679. 195. Açikgöz A, Açikgöz G, Tunga U, Otan F (2006) Characteristics and prevalence of non-syndrome multiple supernumerary teeth: a retrospective study. Dentomaxillofacial Radiol 35(3):185-190. 196. Nagaveni NB, Umashankara KV, Radhika NB, Praveen RP, Manjunath S (2010) Maxillary paramolar: report of a case and literature review. Arch Orofacial Sci 5(1):24-28. 197. Nayak G, Shetty S, SinghI, Pitalia D (2012) Paramolar – A supernumerary molar: A case report. Dent Res J (Isfahan) 9(6):797-803. PMID: 23559961. 198. Ungar PS, Grine FE, Teaford MF, Pérez-Pérez A (2001) A review of interproximal wear grooves on fossil hominin teeth with new evidence from Olduvai Gorge. Arch Oral Biol 46:285–292. doi.10.1016/S0003- 9969(00)00128-X. 199. Chavaillon J, Piperno M eds (2004) Studies on the Early Paleolithic site of Melka Kunture, Ethiopia. (Istituto Italiano di Preistoria e Protostoria, Florence). 200. Piperno M, Bulgarelli GM (2004) Prehistoric archaeology. The site of Garba IV. Excavations 1973-1982. Studies on the Early Paleolithic site of Melka Kunture, Ethiopia, eds Chavaillon J, Piperno M (Istituto Italiano di Preistoria e Protostoria, Florence), pp 449-458. 201. Piperno M, Bulgarelli GM, Gallotti R (2004) Prehistoric archaeology. The site of Garba IV. The test trenches A and B and the sounding in Level E. Studies on the Early Paleolithic site of Melka Kunture, Ethiopia, eds Chavaillon J, Piperno M (Istituto Italiano di Preistoria e Protostoria, Florence), pp 581-587. 202. Condemi S (2004) The Garba IV E mandible. Studies on the Early Paleolithic site of Melka Kunture, Ethiopia, eds Chavaillon J, Piperno M (Istituto Italiano di Preistoria e Protostoria, Florence), pp 687-701. 203. Zilberman U, Smith P, Piperno M, Condemi S (2004) Evidence of amelogenesis imperfecta in an early African Homo erectus. J Hum Evol 46:647-653. doi.10.1016/j.jhevol.2004.02.005. 204. Dean MC et al. (2001) Growth processes in teeth distinguish modern humans from Homo erectus and earlier hominins. Nature 414:628-631. 205. Zilberman U, Smith P, Condemi S (2004) Evidence for a genetic disorder affecting tooth formation in the Garba IV child. Studies on the Early Paleolithic site of Melka Kunture, Ethiopia, eds Chavaillon J, Piperno M (Istituto Italiano di Preistoria e Protostoria, Florence), pp 703-713. 206. Aldred MJ, Savarirayan R, Crawford PJM (2003) Amelogenesis imperfecta: a classification and catalogue for the 21st century. Oral Dis 9(1):19-23. doi.10.1034/j.1601.0825.2003.00843.x. 207. Gadhia K, McDonald S, Arkutu N, Malik K (2012) Amelogenesis imperfecta: an introduction. Br Dent J 212(8):377-379. doi.10.1038/sj.bdj.2012.314. 208. Crawford PJM, Aldred M, Bloch-Zupan A (2007) Amelogenesis imperfecta. Orphanet J Rare Dis 2:17. doi.10.1186/1750-1172-2-17. 209. de Lumley MA (2018) Les dents de l’homme du Lazaret. Les Restes Humains Fossiles de la Grotte du Lazaret, Nice, Alpes-Maritimes, ed de Lumley MA (CNRS, Paris), pp 321-384. 210. de Lumley MA (ed) (2018) Les Restes Humains Fossiles de la Grotte du Lazaret, Nice, Alpes-Maritimes. (CNRS, Paris). 211. de Lumley H, Khatib S (2018) Localisation stratigraphique des restes humains fossiles de la Grotte du Lazaret. Les Restes Humains Fossiles de la Grotte du Lazaret, Nice, Alpes-Maritimes, ed de Lumley MA (CNRS, Paris). pp 63-65. 212. Levitan ME, Himel VT (2006) Dens evaginatus: Literature review, pathophysiology, and comprehensive treatment regimen. J Endodont 32(1):1-9. doi.10.1016/j.joen.2005.10.009. 213. Yip WK (1974) The prevalence of dens evaginatus. Oral Surg Oral Med Oral Pathol 38:80–87. 214. Kocsis GS, Marcsik A, Kókai EL, Kocsis KS (2002) Supernumerary occlusal cusps on permanent human teeth. Acta Biol Szegediensis 46(1-2):71-82. 215. Lacy SA, Wu XJ, Trinkaus E (2012) Dentoalveolar paleopathology of the early modern humans from Zhirendong, south China. Int J Paleopathol 2:10-18. doi.10.1016/j.ijpp.2012.06.003. 216. Aguiló L, Gandia J., Cibrian R, Catala M (1999) Primary double teeth. A retrospective clinical study of their morphological characteristics and associated anomalies. Int J Paediatr Dent 9(3):175-183. 217. de Sigueira VCF, Braga TL, Martins MAT, Raitz R, Martins MD (2004) Dental fusion and dens evaginatus in the permanent dentition: Literature review and clinical case report with conservative treatment. J Dent Children 71(1):69-72.

63

218. Anantharaj A, Sudhir R, Praveen P, Karthik Venkataraghavan, Prathibha Rani (2010) Fused supplemental premolars – a case report. SRM Univ J Dent Sci 1(3):259-261. 219. Smith MO, Wojcinski MC (2011) Anomalous double-crowned primary teeth from Pre-Columbian Tennessee: A meta-analysis of hunter–gatherer and agriculturalist samples. Int J Paleopathol 1(3):173-183. doi.10.1016/j.ijpp.2012.02.008. 220. Filhol MH (1889) Note sur une mâchoire humaine trouvée dans la caverne de Malarnaud près de Montseron (Ariège). Bull Soc Philomatique Paris 1:69–82. 221. Boule M (1921) Les Hommes Fossiles. (Masson, Paris). 222. Heim JL, Granat J (1995) La mandibule de l’enfant néandertalien de Malarnaud (Ariège). Anthropol Préhist 106:75-96. 223. Tillier AM, Kaffe I, Arensburg B, Chech M (1998) Hypodontia of permanent teeth among Middle Palaeolithic hominids: An early case dated to ca. 92 000 ± 5000 years BP at the Qafzeh site. Int J Osteoarchaeol 8:1-6. 224. Tompkins RL (1996) Relative dental development of Upper Pleistocene hominids compared to human population variation. Am J Phys Anthropol 99:103–118. 225. Liu W, Zeng XG (1996) The degeneration of third molars and its significance to human evolution (in Chinese with English abstract). Acta Anthropol Sinica 15(3):185-199. 226. Suwa G, White TD, Howell FC (1996) Mandibular postcanine dentition from the Shungura Formation, Ethiopia: Crown morphology, taxonomic allocations, and Plio-Pleistocene hominid evolution. Am J Phys Anthropol 101:247-282. 227. Polder BJ, Van’t Hot MA, Van der Linden FPGM, Kuijpers-Jagtman AM (2004) A meta-analysis of the prevalence of dental agenesis of permanent teeth. Comm Dentist Oral Epidemiol 32:217-226. doi.10.1111/j.1600-0528.2004.00158.x 228. Trinkaus E, Constantin S, Zilhão J eds (2013) Life and Death at the Peştera cu Oase. A Setting for Modern Human Emergence in Europe. (Oxford University Press, New York). 229. Fu QM et al. (2015) An early modern human from Romania with a recent Neanderthal ancestor. Nature 524(7564):216-219. doi.10.1038/nature14558. 230. Reich D et al. (2010) Genetic history of an archaic hominin group from Denisova Cave in Siberia. Nature 468:1053-1060. 231. Mednikova MB (2011) A proximal pedal phalanx of a Paleolithic hominin from Denisova Cave, Altai. Archaeol Ethnol Anthropol Eurasia 39(1):129-138. 232. Mednikova MB (2013) Distal phalanx of the hand of Homo from Denisova Cave stratum 12: A tentative description. Archaeol Ethnol Anthropol Eurasia 54(2):147-155. 233. Sawyer S et al. (2015) Nuclear and mitochondrial DNA sequences from two Denisovan individuals. Proc Natl Acad Sci USA 112(51):15696-15700. doi.10.1073/pnas.1519905112. 234. Hublin JJ, McPherron SP eds (2012) Modern Origins: A North African Perspective. (Springer, Dordrecht). 235. Trinkaus E, Bailey SE, Rougier H (2013) The dental and alveolar remains of Oase 1 and 2. Life and Death at the Peştera cu Oase. A Setting for Modern Human Emergence in Europe, eds Trinkaus E, Constantin S, Zilhão J (Oxford University Press, New York), pp 348-374. 236. Wood B (1991) Koobi Fora Research Project 4: Hominid Cranial Remains. (Oxford University Press, Oxford UK). 237. Glantz M et al. (2008) New hominin remains from . J Hum Evol 55:223-237. doi.10.1016/j.jhevol.2007.12.007. 238. Bailey S, Glantz M, Weaver TD, Viola B (2008) The affinity of the dental remains from Obi-Rakhmat Grotto, Uzbekistan. J Hum Evol 55:238-248. doi.10.1016/j.jhevol.2008.03.004. 239. Hublin JJ et al. (2012) Dental evidence from the Aterian human populations of Morocco. Modern Origins: A North African Perspective, eds Hublin JJ, McPherron SP (Springer, Dordrecht), pp 189-204. doi.10.1007/978-94-007-2929-2_13. 240. Movius HL Jr ed (1975) Excavation of the Abri Pataud. Les Eyzies (Dordogne). Bull Am Sch Prehist Res 30. 241. Nespoulet R, Chiotti L, Henry-Gambier D eds (2013) Le Gravettien final de l’Abri Pataud (Dordogne, France). Fouilles et études 2005-2009. Brit Archaeol Rep Int Series 2458. 242. Billy G (1975) Étude anthropologique des restes humains de l'Abri Pataud, in Excavation of the Abri Pataud. Les Eyzies (Dordogne). Am Sch Prehist Res Bull 30:201-261.

64

243. Villotte S, Chiotti L, Nespoulet R, Henry-Gambier D (2015) Étude anthropologique des vestiges humains récemment découverts issus de la couche 2 de l’Abri Pataud (Les Eyzies-de-Tayac-Sireuil, Dordogne, France). Bull Mém Soc Anthropol Paris 27:158-188. doi.10.1007/s13219-015-0128-3. 244. Henry-Gambier D, Villotte S, Beauval C, Brůžek J, Grimaud-Hervé D (2013) Les vestiges humains: un assemblage original. Le Gravettien final de l’Abri Pataud (Dordogne, France). Fouilles et études 2005- 2009, eds Nespoulet R, Chiotti L, Henry-Gambier D, Br Archaeol Rep Int Series 2458:135-199. 245. Legoux P (1972-74) Étude odontologique des restes humains périgordiens et protomagdaléniens de l'Abri Pataud (Dordogne). Bull Mém Soc Anthropol Paris, Série 12, 9:293-330; Série 13, 1:45-84. 246. Legoux P (1975) Présentation des dents des restes humains de l'Abri Pataud. Bull Am Sch Prehist Res 30:262-304. 247. Grimanis GA, Kyriakides AT, Spyropoulos ND (1991) A survey of supernumerary molars. Quintessence Int 22(12):989-995. 248. Timocin N, Yalcin S, Ozgen M, Tanyeri H (1994) Supernumerary molars and paramolars. J Nihon Univ Sch Dentist 36(2):145-150. 249. Parolia A, Kundabala M, Dahal M, Mohan M, Thomas MS (2011) Management of supernumerary teeth. J Conserv Dentist 14(3):221-224. doi. 10.4103/0972-0707.85791. 250. Loh FC, Yeo JF. (1993) Paramolar with bifid crown. Oral Surg Oral Med Oral Pathol Oral Radiol 76:257– 258. 251. Dhull KS, Dhull RS, Panda S, Acharya S, Yadav S, Mohanty G (2014) Bilateral mandibular paramolars. Int J Clin Pediatr Dent 7(1):40-42. doi.10.5005/jp-journals-10005-1231. 252. Jacobs R, Willems G (2003) Inverted eruption of a supplemental lower premolar: report of an unusual case. Int J Paediatr Dent 13L:46-50. 253. Mohan S, Kankariya H, Fauzdar S (2012) Impacted inverted teeth with their possible treatment protocols. J Maxillofacial Oral Surg.11(4):455-457. doi.10.1007/s12663-011-0307-9. 254. Kuba Y, Taniguchi K, Kuboyama H, Motokawa W, Braham RL (1995) Inverted impaction of second premolar: two case reports. J Clinical Pediatr Dent 19:205–209. 255. Ulusoy AT, Akkacaoglu M, Akan S, Kocadereli I (2009) Reimplantation of an inverted maxillary premolar: Case report of a multidisciplinary treatment approach. J Clin Pediatr Dentist 33(4):279-282. 256. Lalueza-Fox C et al. (2011) Genetic evidence for patrilocal mating behaviour among Neandertal groups. Proc Natl Acad Sci USA 108:250-253. doi.10.1073/pnas.1011553108. 257. Rosas et al. (2012) Les Néandertaliens d’El Sidrón (Asturies, Espagne). Actualisation d’un nouvel échantillon. L’Anthropol 116(1):57-76. doi.10.1016/j.anthro.2011.12.003. 258. Cañaveras JC et al. (2018) Estudio geoarqueológico de la cueva de El Sidrón (Piloña, Asturias). Bol Geol Min 129(1/2):107-128. doi.10.21701/bolgeomin.129.1.005. 259. Dean MC et al. (2013) Longstanding dental pathology in Neandertals from El Sidrón (Asturias, Spain) with a probable familial basis. J Hum Evol 64:678-686. doi.10.1016/j.jhevol.2013.03.004. 260. Bartucz L et al. (1940) Die Mussolini-Höhle (Subalyuk) bei Cserépfalu. Geol Hung Ser Palaeontol 14:1- 352. 261. Bartucz L (1940) Der Urmensch der Mussolini-Höhle. Geol Hung Ser Palaeontol 14:49-105. 262. Thoma A (1963) The dentition of the Subalyuk Neandertal child. Z Morphol Anthropol 54(2):127-150. 263. Bruszt P (1953) A felső tejszemfogak kettőzöttségéről. Fogorvosi Szemle 7:1-4. 264. Tannenbaum KA, Alling EE (1963) Anomalous tooth development. Case reports of gemination and twinning. Oral Surg Oral Med Oral Pathol 16(7):883-887. 265. Duncan WK, Helpin ML (1987) Bilateral fusion and gemination: A literature analysis and case report. Oral Surg Oral Med Oral Path 64(1):82-87. 266. Taylor GS (1972) Characteristics of supernumerary teeth in the primary and permanent dentition. Dent Pract Dent Rec 22:203-208. 267. Tasa GL, Lukacs JR (2001) The prevalence and expression of primary double teeth in western . J Dentist Child 68(3):196-200. 268. Liu W et al. (2010) Human remains from Zhirendong, south China, and modern human emergence in east Asia. Proc Natl Acad Sci USA 107 (45):19201-19206. doi.10.1073/pnas.1014386107. 269. Cai YJ et al. (2017) The age of human remains and associated fauna from the Zhiren Cave in , southern China. Quatern Int 434:84-91. doi.10.1016/j.quaint.2015.12.088. 270. Hillson SW (1996) Dental Anthropology. (Cambridge University Press, Cambridge UK). 271. Escobar V, Melnick M, Conneally PM (1976) The inheritance of bilateral rotation of maxillary central incisors. Am J Phys Anthropol 45:109-116.

65

272. Ling JYK, Wong RWK (2010) Incisor winging in Chinese. Open Anthropol J 3:8-11. 273. Uyeno DS, Lugo A (1996) Dens evatinatus: a review. J Dentist Child 63:328–332. 274. Li ZY et al. (2017) Late Pleistocene archaic human crania from Xuchang, China. Science 355:969-972. doi.10.1126/science.aal2482. 275. Webb PAO, Suchey JM (1985) Epiphyseal union of the anterior iliac crest and medial clavicle in a modern multiracial sample of American males and females. Am J Phys Anthropol 68:457-466. 276. Formicola V, Frayer DW, Heller JA (1990) Bilateral absence of the lesser trochanter in a Late Epigravettian skeleton from Arene Candide (Italy). Am J Phys Anthropol 83:425-437. 277. Belcastro MG, Rastelli E, Mariotti V (2008) Variation in the degree of sacral vertebral body fusion in adulthood in two European modern skeleton collections. Am J Phys Anthropol 135:149-160. doi.org/10.1002/ajpa.20716. 278. Bonmatí A et al. (2010) Middle Pleistocene lower back and pelvis from an aged human individual from the Sima de los Huesos site, Spain. Proc Natl Acad Sci USA 107(43):18386-18391. doi/10.1073/pnas.1012131107. 279. Willis TA (1931) The separate neural arch. J Bone Joint Surg 13(4):709-721. 280. Roche MB, Rowe GG (1951) The incidence of separate neural arch and coincident bone variations. Anat Rec 109:233-252. 281. Fredrickson BE, Baker D, McHolick WJ, Yuan HA, Lubicky JP (1984) The natural history of spondylolysis and spondylolisthesis. J Bone Joint Surg 66-A(5):699-707. 282. Sakai T, Sairyo K, Suzue N, Kosaka H, Yasui N (2010) Incidence and etiology of lumbar spondylolysis: review of the literature. J Orthop Sci 15:281-288. doi.10.1007/s00776-010-1454-4. 283. Merbs CF (1989) Spondylolysis: its nature and anthropological significance. Int J Anthropol 4(3):163-169. 284. McTimoney M, Micheli LJ (2003) Current evaluation and management of spondylolysis and spondylolisthesis. Curr Sports Med Rep 2:41–46. 285. Crawford CH et al. (2015) Current evidence regarding the etiology, natural history, and prognosis of pediatric lumbar spondylolysis: A report from the Scoliosis Research Society Evidence-based Medicine Committee. Spine Deformity 3:12-29. doi.10.1016/j.jspd.2014.06.005. 286. Rivière E (1887) De l’Antiquité de l’Homme dans Alpes-Maritimes (Librairie J.-B. Baillière et Fils, Paris). 287. Cartailhac E (1912) Les Grottes de Grimaldi (Baoussé-Roussé) II(II): Archéologie. (Imprimérie de Monaco, Monaco). 288. Henry-Gambier D (2001) Le Sépulture des Enfants de Grimaldi (Baoussé-Roussé, Italie). Anthropologie et Palethnologie Funéraire des Populations de la Fin du Paléolithique Supérieur. (Éditions du Comité des travaux historiques et scientifiques. Réunion des Musées Nationaux, Paris). 289. Smith P, Arensburg B (1977) A Mousterian skeleton from Kebara Cave. Eretz-Israel 13:164-176. 290. Bar-Yosef O, Vandermeersch B (eds) (1991) Le Squelette Moustérien de Kébara 2. (CNRS, Paris). 291. Tillier AM, Arensburg B, Vandermeersch B, Chech M (2003) New human remains from Kebara Cave (Mount Carmel). The place of the Kebara hominids in the Levantine Mousterian fossil record. Paléorient 29(2):35-62. 292. Bar-Yosef O et al. (1992) The excavations in Kebara Cave, Mt. Carmel. Curr Anthropol 33(15):497-550. 293. Arensburg B (1991) The vertebral column, thoracic cage, and hyoid bone. Le Squelette Moustérien de Kébara 2, eds Bar Yosef O, Vandermeersch B (CNRS, Paris) pp 113-146. 294. Rak Y (1991). The pelvis. Le Squelette Moustérien de Kébara 2, eds Bar Yosef O, Vandermeersch B (CNRS, Paris) pp 147-156. 295. Cardoso HFV, Pereira V, Rios L (2014) Chronology of fusion of the primary and secondary ossification centers in the human sacrum and age estimation in child and adolescent skeletons. Am J Phys Anthropol 153:214-225. doi.10.1002/ajpa.22422. 296. Brown F, Harris J, Leakey R, Walker A (1985) Early Homo erectus skeleton from west Lake Turkana, . Nature 316:788–792. 297. Walker A, Leakey R, eds (1993) The Nariokotome Homo erectus skeleton. (Harvard University Press, Cambridge MA). 298. Schiess R, Haeusler M (2013) No skeletal dysplasia in the Nariokotome boy KNM-WT 15000 (Homo erectus)—A reassessment of congenital pathologies of the vertebral column. Am J Phys Anthropol 150:365–374. doi.10.1002/ajpa.22211. 299. Haeusler M, Schiess R, Boeni T (2013) Evidence for juvenile disc herniation in a Homo erectus boy skeleton. Spine 38(3):E123–E128. doi.10.1097/BRS.0b013e31827cd245.

66

300. von Lüdinghausen M, Schindler G, Kageyama I, Pomaroli A (2002) The third occipital condyle, a constituent part of a median occipito-atlanto-odontoid joint: a case report. Surg Radiol Anat 24:71–76. doi.10.1007/s00276-002-0018-5. 301. Prescher A, Brors D, Adam G (1996) Anatomic and radiologic appearance of several variants of the craniocervical junction. Skull Base Surg 6(2):83-94. 302. Fojtová M (2007) An analysis of epigenetic traits on the skeletal remains of the Old Slavonic populations from Dolní Věstonice, Czech Republic. Anthropol (Brno) 45(1):81-90. 303. Saralaya VV, Bv M, Vaderav R, Jiji PJ (2012) Occipital condyle morphometry and incidence of condylus tertius: Phylogenetic and clinical significance. Clin terapeut 163(6):479-82. 304. Vallois HV, de Félice S (1977) Les Mésolithiques de France. Arch Inst Paléontol Hum 37:1-194. 305. Dastugue J, de Lumley MA (1976) Les maladies des hommes préhistoriques du Paléolithique et du Mésolithique. La Préhistorie Française 1: Civilisations Paléolithiques et Mésolithiques, ed de Lumley H (CNRS, Paris), pp 612-622. 306. Dastugue J (1979) Pathologie des Mésolithiques de France. L’Anthropol 83(4):602-625. 307. Dastugue J (1977) Rochereil 1 – Pathologie. Arch Inst Paléontol Hum 37:106-107. 308. Auerbach BM, Raxter MH (2008) Patterns of clavicular bilateral asymmetry in relation to the humerus: variation among humans. J Hum Evol 54:663-674. doi.10.1016/j.jhevol.2007.10.002. 309. Cho TJ, Choi IH, Chung CY, Hwang JK (2000) The Sprengel deformity. Morphometric analysis using 3D- CT and its clinical relevance. J Bone Joint Surg 82-B:711-718. 310. Solecki RS (1963) Prehistory in Shanidar Valley, northern . Science 139:179-193. 311. Cowgill LW, Trinkaus E, Zeder MA (2007) Shanidar 10: A Middle Paleolithic immature distal lower limb from Shanidar Cave, Iraqi Kurdistan. J Hum Evol 53:213-223. doi.10.1016/j.jhevol.2007.04.003. 312. Shinohara H (1999) The size and position of the sacral hiatus in man. Okajimas Folia Anat Jap 76(2):89-94. 313. Radlauer C (1908) Beiträge zur Anthropologie des Kreuzbeines. Gegenbaurs Morphol Jahrb 38:323-447. 314. Trotter M, Lanier PF (1945) Hiatus canalis sacralis in American whites and negroes. Hum Biol 17:368-381. 315. Sekiguchi M, Yabuki S, Satoh K, Kikuchi S (2004) An anatomic study of the sacral hiatus: A basis for successful caudal epidural block. Clin J Pain 20(1):51-54. 316. Aggarwal A, Aggarwal A, Harjeet Sahni D (2009) Morphometry of sacral hiatus and its clinical relevance in caudal epidural block. Surg Radiol Anat 31:793. doi.10.1007/s00276-009-0529-4 317. Chhabra N (2014) An anatomical study of the size and position of sacral hiatus: Its importance in caudal epidural block. Int J Health Sci Res 4(12):189-196. 318. Heim JL (1976) Les hommes fossiles de La Ferrassie I: Le gisement. Les squelettes adultes (crâne et squelette du tronc). Arch Inst Paléontol Hum 35:1-331. 319. Solecki RS (1960) Three adult Neanderthal skeletons from Shanidar Cave, northern Iraq. Annu Rep Smithson Inst 1959:603-635. 320. Ogilvie MD, Hilton CE, Ogilvie CD (1998) Lumbar anomalies in the Shanidar 3 Neandertal. J Hum Evol 35:597-610. 321. Franciscus RG, Churchill SE (2002) The costal skeleton of Shanidar 3 and a reappraisal of Neandertal thoracic morphology. J Hum Evol 42:303-356. doi.10.1006/jhev.2001.0528. 322. Gómez-Olivencia A, Couture-Veschambre C, Madelaine S, Maureille B (2013) The vertebral column of the Regourdou 1 Neandertal. J Hum Evol 64:582–607. doi.10.1016/j.jhevol.2013.02.006. 323. Lanier RR Jr (1944) Length of first, twelfth, and accessory ribs in American whites and negroes; their relationship to certain vertebral variations. Am J Phys Anthropol 2(2):137-146. 324. Barnes E (1994) Developmental Defects of the in Paleopathology. (University of Colorado Press, Boulder). 325. Allbrook DB (1955) The East African vertebral column. Am J Phys Anthropol 13:489–514. 326. Bagnall KM, Harris PF, Jones PRM (1984) A radiographic study of variations in the human fetal spine. Anat Rec 208(2):265-270. doi.10.1002/ar.1092080213, 327. Steele DG, Bramblett CA (1988) The Anatomy and Biology of the Human Skeleton. (Texas A&M University Press, College Station). 328. Merks HM et al. (2005) Prevalence of rib anomalies in normal Caucasian children and childhood cancer patients. Europ J Med Gen 48(2):113-119. doi.10.1016/j.ejmg.2005.01.029. 329. Hasabe K (1912) Die Wirbelsäule der Japaner. Z Morphol Anthropol 15:259-380. 330. Holliday TW (1995) Body Size and Proportions in the Late Pleistocene Western Old World and the Origins of Modern Humans. PhD Thesis, University of New .

67

331. Lanier RR Jr (1939) The presacral vertebrae of American white and negro males. Am J Phys Anthropol 25(3):341-420. 332. Davies KM, Recker RR, Heaney RP (1989) Normal vertebral dimensions and normal variation in serial measurements of vertebrae. J Bone Min Res 4(3):341-349. 333. Scoles PV, Latimer BM, DiGiovanni BF, Vargo E, Bauza S, Jellema LM (1991) Vertebral alterations in Scheuermann’s kyphosis. Spine 16(5):509-515. 334. Shi C, Nichizawa S, Adachi K, Kimura T (1999) Morphological comparison of the vertebrae in humans and some other mammals. Anthropol Sci 107(1):3-19. 335. Holliday TW, Hilton CE (2010) Body proportions of circumpolar peoples as evidenced from skeletal data: Ipiutak and Tigara (Point Hope) versus Kodiak Island Inuit. Am J Phys Anthropol 142:287-302. doi.10.1002/ajpa.21226. 336. Stoddard A, Osborn JF (1979) Scheuermann’s disease or spinal osteochondrosis. Its frequency and relationship with . J Bone Joint Surg 61B(1):56- 58. 337. Ali RM, Green DW, Patel TC (2000) Scheuermann’s kyphosis. Curr Opin Orthop 11:131-136. 338. Palazzo C, Sailhan F, Revel M (2014) Scheuermann’s disease: an update. Joint Bone Spine 81:209-214. doi.10.1016/j.jbspin.2013.11.012. 339. Makurthou AA et al. (2013) Scheuermann disease. Evaluation of radiological criteria and population prevalence. Spine 38(19):1690-1694. doi.10.1097/BRS.0b013e31829ee8b7. 340. Ríos L et al. (2015) Possible further evidence of low genetic diversity in the El Sidrón (Asturias, Spain) Neandertal group: congenital clefts of the atlas. PLoS One 10(9):e0136550. doi.10.1371/journal.pone.0136550. 341. Gómez-Olivencia A (2013) The presacral spine of the La Ferrassie 1 Neandertal: a revised inventory. Bull Mém Soc Anthropol Paris 25:19-38. doi.10.1007/s13219-012-0064-4. 342. Ferembach D (1962) La Nécropole Épipaléolithique de Taforalt (Maroc oriental). Étude des Squelettes Humains (Mission Universitaire et Culturelle Française au Maroc, Rabat). 343. Roche J (1963) L’Épipaléolithique Marocain. (Didier, Paris). 344. Humphrey LT et al. (2014) First evidence for caries and exploitation of starchy plant foods in Pleistocene hunter-gatherers from Morocco. Proc Natl Acad Sci USA 111(3):945-959. doi.10.1073/pnas.1318176111. 345. Ferembach D (1963) Frequency of spina bifida occulta in prehistoric human skeletons. Nature 199:100- 101. 346. Saluja PG (1988) The incidence of spina bifida occulta in a historic and a modern London population. J Anat 158:91-93. 347. Eubanks JD, Cheruvu VK (2009) Prevalence of sacral spina bifida occulta and its relationship to age, sex, race, and the sacral table angle. An anatomic, osteologic study of three thousand one hundred specimens. Spine 34(15):1539-1543. 348. Fidas A, MacDonald HL, Elton RA, Wild SR, Chisholm GD, Scott R (1987) Prevalence and patterns of spina bifida occulta in 2707 norml adults. Clin Radiol 38:537-542. 349. Broglio A (2000) The Epigravettian funeral structure of the Villabruna shelter A: a tendency to mythicize the dead? Étud Rech Archéol Univ Liège 102:121-129. 350. Vercellotti G, Alciati G, Richards MP, Formicola V (2008) The Late Upper Paleolithic skeleton Villabruna 1 (Italy): a source of data on biology and behavior of a 14,000 year-old hunter. J Anthropol Sci 86:143-163. 351. Oxilia G et al. (2015) Earliest evidence of dental caries manipulation in the Late Upper Palaeolithic. Sci Rep 5:12150. doi.10.1038/srep12150. 352. Nicoll EA (1949) Fractures of the dorso-lumbar spine. J Bone Joint Surg 31-B:376-394. 353. Magerl F, Aebi M, Gertzbein SD, Harms J, Nazarian S (1994) A comprehensive classification of thoracic and lumbar injuries. Europ Spine J 3:184-201. 354. Verneau R (1906) Les Grottes de Grimaldi (Baoussé-Roussé) II(I). Anthropologie. (Imprimérie de Monaco, Monaco). 355. Mussi M (1986) Italian Upper Palaeolithic and Mesolithic burials. Hum Evol 1(6):545-556. 356. Mussi M (1986) On the chronology of the burials found in the Grimaldi . Antropol Contemp 9:95- 104. 357. Formicola V (1990) The triplex burial of Barma Grande (Grimaldi, Italy). Homo 39:130- 143 358. Formicola V, Pettitt PB, Del Lucchese A (2004) A direct radiocarbon AMS date on the Barma Grande 6 Upper Paleolithic skeleton. Curr Anthropol 45(1):114-117. 359. Churchill SE, Fromicola V (1997) A case of marked bilateral asymmetry in the upper limbs of an Upper Palaeolithic male from Barma Grande (Liguria), Italy. Int J Osteoarchaeol 7:18-38.

68

360. Pearson OM, Lieberman DE (2004) The aging of Wolff’s “law”: ontogeny and responses to mechanical loading in cortical bone. Yrbk Phys Anthropol 47:63–99. doi.10.1002/ajpa.20155. 361. Trinkaus E, Buzhilova AP (2012) The death and burial of Sunghir 1. Int J Osteoarchaeol 22:655-666. doi.10.1002/oa.1227. 362. Kozlovskaya MV, Mednikova MB (2000) Catalogue of photos and tables on materials from Sunghirian graves 1 and 2. Homo Sungirensis: Upper Palaeolithic Man: Ecological and Evolutionary Aspects of the Investigation, eds Alexeeva TI et al. (Scientific World, Moscow), pp 85–144. 363. Trinkaus E, Holliday TW, Auerbach BM (2014) Neandertal clavicle length. Proc Natl Acad Sci USA 111:4438-4442. doi.10.1073/pnas.1402439111. 364. Edelson JG, Zuckerman J, Hershkovitz I (1993) Os acromiale: anatomy and surgical implications. J Bone Joint Surg 75-B:551-555. 365. Burbank KM, Lemos MJ, Bell G, Lemos DW (2007) Incidence of os acromiale in patients with shoulder pain. Am J Orthop 36(3):153-155. 366. Nicholson GP, Goodman DA, Flatow EL, Bigliani LU (1996) The acromion: morphologic condition and age-related changes. A study of 420 scapulas. J Shoulder Elbow Surg 5(1):1-11. doi.10.1016/S1058- 2746(96)80024-3. 367. Sammarco VJ (2000) Os acromiale: frequency, anatomy, and clinical implications. J Bone Joint Surg 82- A(3):394-400. 368. Liberson F (1937) Os acromiale – a contested anomaly. J Bone Joint Surg 19-A(3):683-689. 369. Verna C, Villotte S (2016) Several infra-cranial non-metric variations in a single Neolithic individual: A common genetic underlying predisposition? Int J Osteoarchaeol 26:1113-1118. doi.10.1002/oa.2522. 370. James SLJ, Davies AM (2006) Atraumatic avulsion of the lesser trochanter as an indicator of tumour infiltration. Europ Radiol 16:512-514. doi.10.1007/s00330-005-2675-3. 371. Hamanishi C (1980) Congenital short femur. Clinical, genetic and epidemiological comparison of the naturally occurring condition with that caused by thalidomide. J Bone Joint Surg 62-B(3):307-320. 372 Conroy GC, Pickford M, Senut B, Mein P (1993) Diamonds in the desert: the discovery of Otavipithecus namibiensis. Evol Anthropol 2:46–52. 373 Grine FE, Jungers WL, Tobias PV, Pearson OM (1995) Fossil Homo femur from Berg Aukas, northern Namibia. Am J Phys Anthropol 97:151–185. 374 Trinkaus E, Ruff CB, Conroy GC (1999) The anomalous archaic Homo femur from Berg Aukas, Namibia: A biomechanical assessment. Am J Phys Anthropol 110:379-391. doi.10.1002/(SICI)1096- 8644(199911)110:3<379::AID-AJPA9>3.0.CO;2-J 375 Ruff CB, Burgess ML, Squyres N, Junno JA, Trinkaus E (2018) Lower limb articular scaling and body mass estimation in Pliocene and Pleistocene humans. J Hum Evol 115: 85-111. doi.10.1016/j.jhevol.2017.10.014. 376 Shackelford LL (2005) Patterns of Geographic Variation in the Postcranial Robusticity of Late Upper Paleolithic Humans. Ph.D. Thesis, Washington University. 377 Rodríguez L, Carretero JM, García-González R, Arsuaga JL (2018) Cross-sectional properties of the lower limb long bones in the Middle Pleistocene Sima de los Huesos sample (Sierra de Atapuerca, Spain). J. Hum Evol 117:1-12. doi.10.1016/j.jhevol.2017.007. 378 Trinkaus E (2018) Epipaleolithic human appendicular remains from Ein Gev I, Israel. C R Palevol 17:616- 627. doi.10.1016/j.crpv.2018.03.002. 379 Trinkaus E, Ruff CB (2012) Femoral and tibial diaphyseal cross-sectional geometry in Pleistocene Homo. PaleoAnthropology 2012:13-62. doi.10.4207/PA.2012.ART69. 380 Trinkaus E (1993) Femoral neck-shaft angles of the Qafzeh-Skhul early modern humans, and activity levels among immature Near Eastern Middle Paleolithic hominids. J Hum Evol 25:393-416. doi.10.1006/jhev.1993.1058. 381 Gilligan I, Chandraphak S, Mahakkanukrauh P (2013) Femoral neck-shaft angle in humans: variation relating to climate, clothing, lifestyle, sex, age and side. J Anat 223:133-151. doi.10.1111/joa.12073. 382. Vermeersch PM ed (2002) Palaeolithic Quarrying Sites in Upper and Middle Egypt. (Leuven University Press, Leuven). 383. Thoma A (1984) Morphology and affinities of the Nazlet Khater man. J Hum Evol 13:287-296. 384. Vermeersch PM (2002). Two Upper Palaeolithic burials at Nazlet Khater. Palaeolithic Quarrying Sites in Upper and Middle Egypt, ed Vermeersch PM (Leuven University Press, Leuven), pp 273-282. 385. Crevecoeur I, Villotte S (2006) Atteintes pathologiques de Nazlet Khater 2 et activité minière au début du Paléolithique supérieur en Egypte. Bull Mém Soc Anthropol Paris 18(3-4):165-175.

69

386. Piveteau J (1959) Les restes humains de la grotte du Regordou (Dordogne). C R Acad Sci Paris 248D:40– 44. 387. Madelaine S et al. (2008) Nouveaux restes humains moustériens rapportés au squelette Néandertalien de Regourdou 1 (Regourdou, commune de Montignac, Dordogne, France). Paleo 20:101-114. 388. Maureille B, Gómez-Olivencia A, Couture-Veschambre C, Madelaine S, Holliday T (2015) Nouveaux restes humains provenant du gisement de Regourdou (Montignac-sur-Vézère, Dordogne, France). Paléo 26:117–138. 389. Maureille B et al. (2016) New data on the possible Neandertal burial at Regourdou (Montignac-sur-Vézère, Dordogne, France). Qu’est-ce Qu’Une Sépulture? eds Lauwers M, Zemour A (Éditions APDCA, Antibes), pp 175-191. 390. Coutinho Nogueira D, Santos F, Courtaud P, Couture-Veschambre C (2017) Le calcanéus “Regourdou 2”: étude morphométrique comparative et discussion autour de sa place dans la variabilité des Néandertaliens. Paleo 28:71-89. 391. Vandermeersch B, Trinkaus E (1995) The postcranial remains of the Régourdou 1 Neandertal. The shoulder and arm remains. J Hum Evol 28:439–476. 392. Volpato V, Macchiarelli R, Guatelli-Steinberg D, Fiore I, Bondioli L, Frayer DW (2012) Hand to mouth in a Neandertal: Right-handedness in Regourdou 1. PLoS One 7:e43949. doi.10.1371/journal.pone.0043949. 393. Meyer V (2013) Apport de la Reconstruction Virtuelle du Bassin de Regourdou 1 (Dordogne, France) à la Connaissance de l’Obstétrique Néandertalienne. Thèse de Doctorat, Université Bordeaux 1. 394. Meyer V, Brůžek J, Couture C, Madelaine S, Maureille B (2011) Un nouveau bassin Néandertalien: description morphologique des restes pelviens de Regourdou 1 (Montignac, Dordogne, France). Paleo 22:207-222. 395. Plochocki JH (2002) Directional bilateral asymmetry in human sacral morphology. Int J Osteoarchaeol 12:349-355. doi.10.1002/oa.633. 396. Boulay C et al. (2006) Three-dimensional study of pelvic asymmetry on anatomical specimens and its clinical perspectives. J Anat 208:21-33. 397. Kurki HK (2017) Bilateral asymmetry in the human pelvis. Anat Rec 300:653-665. doi.10.1002/ar.23546. 398. Al-Eisa E, Egan D, Deluzio K, Wasserburg R (2006) Effects of pelvic skeletal asymmetry on trunk movement. Spine 31(3):E71-E79. 399. Bukhman AI (1984) Roentgenological studies of the children’s skeletons from the Upper Palaeolithic site Sungir (in Russian). Sungir Anthropological Investigations, eds Zubov AA, Kharitonov VM (Scientific World, Moscow), pp 203–204. 400. Buzhilova AP (2000) The analysis of anomalies and indicators of physiological stress in non-mature Sunghir individuals. Homo Sungirensis. Upper Palaeolithic Man: Ecological and Evolutionary Aspects of the Investigation, eds Alexeeva TI et al. (Scientific World, Moscow), pp 302-314. 401. Formicola V, Buzhilova AP (2004) Double child burial from Sunghir (Russia): Pathology and inferences for Upper Paleolithic funerary practices. Am J Phys Anthropol 124:189–198. doi.10.1002/ajpa.10273. 402. Shackelford LL, Trinkaus E (2002) Late Pleistocene human femoral diaphyseal curvature. Am J Phys Anthropol 118:359-370. doi.10.1002/ajpa.10093. 403. Tardieu C, Trinkaus E (1994) The early ontogeny of the human femoral bicondylar angle. Am J Phys Anthropol 95:183-195. doi.10.1002/ajpa.1330950206. 404. Tong H, Shang H, Zhang S, Chen F (2004) A preliminary report on the newly found Tianyuan Cave, a Late Pleistocene human fossil site near Zhoukoudian. Chin Sci Bull 49:853-857. doi.10.1360/03wd0561. 405. Shang H, Tong H, Zhang S, Chen F, Trinkaus E (2007) An early modern human from Tianyuan Cave, Zhoukoudian, China. Proc Natl Acad Sci USA 104:6573-6578. doi.10.1073/pnas.0702169104. 406. Fernández-Jalvo Y, Andrews P (2010) Taphonomy of the human remains from Tianyuandong. The Early Modern Human from Tianyuan Cave, China, eds Shang H, Trinkaus E (Texas A&M University Press, College Station), pp 205-210. 407. Villotte S, Samsel M, Sparacello V (2017) The paleobiology of two adult skeletons from the Baousso da Torre (Bausu da Ture) (Liguria, Italy): Implications for Gravettian lifestyle. C R Palevol 16:462-473. doi.10.1016/j.crpv.2016.09.004. 408. Villotte S, Henry-Gambier D (2010) The rediscovery of two Upper Palaeolithic skeletons from Baousso da Torre cave (Liguria, Italy). Am J Phys Anthropol 141:3-6. 409. Villotte S, Knüsel CJ, Mitchell, PD, Henry-Gambier D (2011) Probable carpometacarpal and tarsal coalition from the Baousso da Torre cave (Italy): Implications for burial selection during the Gravettian. J Hum Evol 61:117-120. doi.10.1016/j.jhevol.2011.01.007.

70

410. Case DT, Burnett SE (2012) Identification of tarsal coalition and frequency estimates from skeletal samples. Int J Osteoarchaeol 22:667-684. doi.10.1002/oa.1228. 411. Alemohammad AM, Nakamura K, El-Sheneway M, Viegas SF (2009) Incidence of carpal boss and osseous coalition: an anatomic study. J Hand Surg 34A:1-6. doi.10.1016/j.jhsa.2008.08.025. 412. Porrino J, Maloney E, Chew FS (2015) Current concepts of the carpal boss: Pathophysiology, symptoms, clinical or imaging diagnosis, and management. Curr Prob Diag Radiol 44(5):462-468. 413. Trinkaus E (2006) The upper limb remains. Early Modern Human Evolution in Central Europe: The People of Dolní Vĕstonice and Pavlov, eds Trinkaus E, Svoboda JA (Oxford University Press, New York), pp 327-372. 414. Shaw CN (2011) Is ‘hand preference’ coded in the hominin skeleton? An in-vivo study of bilateral morphological variation. J Hum Evol 61:480-487. doi.10.1016/j.jhevol.2011.06.004. 415. Haugen IK et al. (2011) Prevalance, incidence and progression of hand osteoarthritis in the general population: the Framingham Osteoarthritis Study. Ann Rheum Dis 70(9):1-15. doi.10.1136/ard.2011.150078. 416. Weiss KE, Rodner CM (2007) Osteoarthritis of the wrist. J Hand Surg 32(5):725-746. 417. Kivell TL, Begun DR (2007) Frequency and timing of scaphoid-centrale fusion in hominoids. J Hum Evol 52(3):321-340. doi.10.1016/j.jhevol.2006.10.002. 418. Heim JL (1982) Les hommes fossiles de La Ferrassie II: Les squelettes adultes (squelettes des membres). Arch Inst Paléontol Hum 38:1-272. 419. Vandermeersch B (1991) La ceinture scapulaire et les membres supérieures. Le Squelette Moustérien de Kébara 2, eds Bar-Yosef, O., Vandermeersch B. (CNRS, Paris), pp 157-178. 420. Solecki RS (1961) New anthropological discoveries at Shanidar, northern Iraq. Trans NY Acad Sci 23:690- 699. 421. Doman AN, Marcus NW (1990) Congenital bipartite scaphoid. J Hand Surg 15A(6):869-873.

71