You are on page 1of 12

Journal of Human Evolution 77 (2014) 155e166

Contents lists available at ScienceDirect

Journal of Human Evolution


journal homepage: www.elsevier.com/locate/jhevol

Food material properties and early hominin processing techniques


Katherine D. Zink a, *, Daniel E. Lieberman a, Peter W. Lucas b, 1
a
Department of Human Evolutionary Biology, Harvard University, 11 Divinity Avenue, Cambridge, MA 02138, USA
b
Department of Anthropology, The George Washington University, 2110 G St NW, Washington, DC 20052, USA

a r t i c l e i n f o a b s t r a c t

Article history: Although early Homo is hypothesized to have used tools more than australopiths to process foods prior to
Received 23 August 2013 consumption, it is unknown how much the food processing techniques they used altered the material
Accepted 21 June 2014 properties of foods, and therefore the masticatory forces they generated, and how well they were able to
Available online 29 October 2014
comminute foods. This study presents experimental data on changes to food material properties caused
by mechanical tenderization (pounding with a stone tool) and cooking (dry roasting) of two foods likely
Keywords:
to have been important components of the hominin diet: meat and tubers. Mechanical tenderization
Diet
significantly decreased tuber toughness by 42%, but had no effect on meat toughness. Roasting signifi-
Food processing
Cooking
cantly decreased several material properties of tubers correlated with masticatory effort including
Tenderizing toughness (49%), fracture stress (28%) and elastic modulus (45%), but increased the toughness (77%),
Early Homo fracture stress (50%e222%), and elastic modulus of muscle fibers in meat (308%). Despite increasing
Oldowan many material properties of meat associated with higher masticatory forces, roasting also decreased
Fire measured energy loss by 28%, which likely makes it easier to chew. These results suggest that the use of
food processing techniques by early Homo probably differed for meat and tubers, but together would
have reduced masticatory effort, helping to relax selection to maintain large, robust faces and large,
thickly enameled teeth.
© 2014 Elsevier Ltd. All rights reserved.

Introduction evidence such as bone cut marks and stone tool remains (Bunn,
1981, 1994, 2007; Bunn and Kroll, 1986; Dominguez-Rodrigo
The evolution of the genus Homo is marked by reduction of the et al., 2002; Plummer, 2004; Dominguez-Rodrigo and Barba,
size of the masticatory apparatus. Compared with gracile austral- 2006). Meat is a high quality food source that is calorically dense,
opiths, early Homo taxa (particularly Homo erectus) possessed less highly digestible, and an important source of protein and fat. From a
robust and buttressed faces, and smaller postcanine teeth (Brace, masticatory perspective, however, consumption of raw meat may
1967; Wolpoff, 1973; Chamberlain and Wood, 1985; McHenry, be a challenge. Muscle tissue comprises elastic contractile fibers
1994; Lahr and Wright, 1996; Wood and Collard, 1999; hierarchically bound by connective tissue. Under compressive,
Lieberman, 2011). Additionally, Eng et al. (2013) have shown that limited-space environments like the area between occluding teeth,
early Homo produced maximum masticatory muscle forces that meat fractures do not effectively propagate. The low-crested
were on average 66% lower than in gracile australopiths. These bunodont molars of apes and hominins appear to be especially
morphological changes are hypothesized to signal reduced masti- poor at fracturing meat, and according to some accounts it takes
catory effort within the genus, and are assumed to have been made chimpanzees 4.0e11.5 h to chew small (~4 kg) animal carcasses
possible by a change in diet to softer foods and/or higher quality, (Goodall, 1986; Wrangham and Conklin-Brittain, 2003). These ob-
energetically dense foods that require fewer chews per calorie servations suggest that increased raw meat consumption by hom-
consumed. inins may have required substantially more chewing effort, which
One often discussed dietary shift is the increased consumption is inconsistent with the relatively smaller, less robust masticatory
of meat by early Homo, which is supported by archaeological apparatus of Homo species.
Another dietary shift that likely evolved in early Homo is
increased reliance on food processing techniques. All human pop-
* Corresponding author. ulations process much of their food before consumption
E-mail addresses: kzink@oeb.harvard.edu, danlieb@fas.harvard.edu (K.D. Zink), (Wrangham and Conklin-Brittain, 2003; Wrangham, 2007). Today
peterwlucas@gmail.com (P.W. Lucas).
1 we fry, boil, bake and steam, and even modern day ‘raw foodists’
Present address: Department of Bioclinical Sciences, Faculty of Dentistry, PO
Box 24923, Safat 13110, Kuwait University, Kuwait. who eschew thermal heating of food (most of which has been

http://dx.doi.org/10.1016/j.jhevol.2014.06.012
0047-2484/© 2014 Elsevier Ltd. All rights reserved.
156 K.D. Zink et al. / Journal of Human Evolution 77 (2014) 155e166

domesticated and is low in fiber and toxins), must expend much 2010; Duan et al., 2010; Dilek et al., 2011). In comparison, less
effort dehydrating, pureeing, blending, juicing and soaking (Baird research has focused on the material property changes in naturally-
and Rodwell, 2005). Although many food processing techniques occurring foods caused by simple cooking or mechanical tenderi-
were developed within the last few thousand years, some of them, zation. Dominy et al. (2008) measured the material property
such as simple mechanical processing and cooking, may have been changes induced by roasting five species of tubers eaten by Hadza
particularly important for early members of the genus Homo. hunter-gatherers, and the effects of boiling, steaming and/or mi-
Tool use and very rudimentary forms of mechanical food pro- crowave cooking is well documented for a limited number of fruits
cessing are not unique to humans. For example, some chimpanzee and vegetables (e.g., Greve et al., 1994a,b; Ng and Waldron, 1997;
populations use stones to pound open hard nuts (Goodall, 1986; Thiel and Donald, 2000; Alvarez and Canet, 2001; Dan et al.,
Boesch and Boesch-Achermann, 2000) and other animals, such as 2003; Lucas, 2004; Beleia et al., 2004a, b).
sea otters, use stones to break open hard mollusk shells (Hall and Because of commercial interest, there is extensive food property
Schaller, 1964). It is therefore possible that the last common literature on a diversity of meat types and processing methods,
ancestor of chimpanzees and hominins also practiced some form of including forms of mechanical (blade) tenderization. However,
rudimentary processing. By the time early Homo evolved, however, most studies measure either Warner-Bratzler ‘shear forces’ (an
mechanical alteration of food likely became much more complex. empirical test measuring maximum fracture force with a notched
Stone tools date to approximately 2.6 mya (millions of years ago) in blade) or perform a texture profile analysis (compressive tests that
the archeological record (Semaw et al., 1997), and may be even assess ‘hardness’, ‘cohesiveness’, ‘springiness’ and ‘chewiness’) as
older (McPherron et al., 2010). Analyses of early Oldowan sites proxies for sensory perception of food by consumers (e.g., Loucks
indicate that hominins used these stones extensively on meat and et al., 1984; Mittal et al., 1992; Combes et al., 2003; King et al.,
plant material (Keeley and Toth, 1981; Semaw et al., 2003; 2003; de Huidobro et al., 2005; Pietrasik and Shand, 2005; Dixon
Dominguez-Rodrigo et al., 2005; Bunn, 2007; Pobiner et al., et al., 2012). These are not true material properties (see below),
2008). Sharp edges on hand axes could have been used to slice meat and therefore have limited utility for modeling food fracture within
and tubers into smaller, more easily ingested particles, while other the oral cavity. A few studies, however, have examined the material
Lower Paleolithic tools including spheroids, hammerstones and properties of cooked meat. The most notable is Purslow (1985),
handaxes could have been used to pound or grind food. These which characterized the fracture properties of beef cooked in a
different kinds of mechanical processing might have significantly water bath. This was followed by a series of papers that analyzed
reduced masticatory effort by reducing ingested particle size and muscle fiber and surrounding connective tissue responses to water-
tenderizing the food. bath cooking (Lewis and Purslow, 1989; Mutungi et al., 1995;
While it is clear then that early Homo had access to mechanical Willems and Purslow, 1997; Christensen et al., 2000, 2004). Un-
food processing technology, the timing of cooking is much more fortunately, all of these studies cooked the meat by boiling it in a
controversial. Wrangham et al. (1999) hypothesized that cooking bag, which precisely controls cooking conditions, but likely causes
softens foods and increases net nutrient availability, helping to substantially less water loss than other more traditional cooking
make possible the evolution of larger brains and body mass com- methods relevant to human evolution.
bined with smaller guts, teeth and less robust faces. Supporting this This study builds on the existing food material property litera-
hypothesis, recent research has demonstrated that cooking signif- ture by measuring the changes in food material properties that
icantly reduces cost of digestion and increases net energy gain in result from using two processing techniques available to early
pythons and mice (Boback et al., 2007; Carmody et al., 2011). A hominins: simple mechanical tenderization (i.e., pounding with a
major problem with ascribing cooking to H. erectus, however, is a stone) and dry roasting. Pounding requires little effort, time or
lack of evidence for controlled fire, let alone cooking use, around manipulative ability and could be easily performed by any hominin
the time of early Homo. The oldest clear evidence of fire in the with a stone. Dry roasting requires no technology other than a fire.
archeological record is from Wonderwerk Cave dated to 1 mya Our experiments focus on tubers and meat because they were likely
(Berna et al., 2012) and Gesher Bonet Ya'aqov at 790 kya (Goren- two important components of hominin diets (e.g., Hatley and
Inbar et al., 2004), but hearths and other features indicative of Kappelman, 1980; Milton, 1999; Laden and Wrangham, 2005;
cooking do not appear until the Middle Paleolithic, leading many Ungar et al., 2006; Bunn, 2007). In addition, these are extremely
researchers to believe that habitual cooking is a relatively recent different foods from a materials standpoint. Raw vegetables such as
behavior (e.g., Brace, 1995; Ragir, 2000; Bunn, 2007; Ben-Dor et al., tubers comprise a latticework of cells under internal turgor pres-
2011; Roebroeks and Villa, 2011). sure and can be modeled as fluid-filled foams (Gibson and Ashby,
Regardless of when cooking and food processing evolved, we do 1997). In contrast, as noted earlier, raw meat is composed of
not know if and by how much early processing techniques could elastic contractile fibers hierarchically bound together with con-
have permitted smaller postcanine dentition and other craniofacial nective tissue. It is therefore reasonable to expect that they will
changes evident in the genus Homo. Therefore, the major goal of respond differently to processing.
this study is to provide experimental data on the material property
changes associated with Lower Paleolithic types of food processing. Food fracture
Material properties describe how a food deforms, and when and
how it will fracture. These intrinsic properties govern the proba- Before discussing the specific hypotheses tested, we first describe
bility of food fracture in the oral cavity and the forces necessary to a model of food fracture in the oral cavity and define the five pa-
create these fractures. rameters that are measured in this study: fracture stress, fracture
Although there is much research devoted to testing food ma- strain, stiffness, toughness, and energy loss. Fig. 1 shows a food item
terial properties, most studies examine the properties of raw foods modeled as a rigid beam between upper and lower teeth (Agrawal
or highly processed foods such as biscuits and cheese, or they et al., 1997). As the teeth come into occlusion, a stress (force per
measure the effects of harvest age, storage, chemical tenderizing, area) is generated within the food, which deforms as it absorbs strain
freeze drying, etc. on properties that relate to the taste/attractive- energy. Initially, the food deforms elastically and will return to its
ness of commercial foods (e.g., Agrawal et al., 1997; Lillford, 2001; original dimensions if the occlusal force is removed (‘elastic’ defor-
Christensen et al., 2003; Beleia et al., 2004a; Goh et al., 2005; Sui mation). As stress increases, however, a yield point is reached and
et al., 2006; Dominy et al., 2008; Vogel et al., 2008; Chang et al., any further deformation becomes permanent (‘plastic’
K.D. Zink et al. / Journal of Human Evolution 77 (2014) 155e166 157

Fracture
Stress
*
Fraccture
Yield Point
*

(Force/Area)
Stress
Slope =
Modulus of Elassticity

S
Strain Fracture
(Displacemennt/Original Lengthh) Strain

Figure 1. Left: A model of food fracture between occluding teeth. Fractures can form as a result of indentation at the cusp tips (solid arrows) or from tension away from the cusps
(hatch-mark arrows). Right: A typical linear stressestrain curve. As the molars bite onto a food item, they create a stress (force/area) that induces the material to strain (proportional
deformation). Continuous application of force creates permanent deformation (yield point) and ultimately leads to the production of a fracture in the material. The slope of the
stressestrain curve is the food's modulus of elasticity, which is a measure of stiffness.

deformation). At some level of stress beyond the yield point, the food more efficient comminution (increased food breakdown per chew),
will fracture (fracture stress). Cracks can be initiated by indentation fewer chews and reduced total masticatory force.
at the cusp tips or by bending stresses distant from the cusps.
Typically, the slope of the elastic portion of the stress-strain
Hypotheses
curve defines the elastic modulus, with higher slopes indicating a
stiffer material. Although the elastic regions of these curves are
Tubers Mechanical tenderization by pounding with an Oldowan
generally linear, many foods have r-shaped stress-strain curves
hammerstone will fracture and damage the internal cellular
where the ‘r’ denotes a stress-strain curve that is concave down-
structure of tubers. Fractures in raw vegetables tend to burst cells,
wards. This curve shape is typical of stiff foods that require
which is facilitated by internal turgor pressure that pre-stresses the
disproportionate stress, compared with strain, at fracture. More
cell walls (Ng and Waldron, 1997; Thiel and Donald, 2000; Lillford,
pliant foods have stress-strain curves that are concave upwards,
2001). The elastic modulus of cell walls exceeds that of intact turgid
giving them a ‘J’ shape (Purslow, 1991a; Sui et al., 2006), and
vegetable tissue by a substantial margin and thus ‘tenderized’
typically require more strain to fracture. This difference is pertinent
vegetables may be stiffer, but the extensive fractures produced by a
to the study of oral food fracture because the limited space between
hammerstone will reduce overall toughness.
occluding teeth means that less stiff foods may not fracture as
In terms of cooking, at temperatures above 40  C pectin sub-
readily. Note that with foods possessing ‘J’ shaped stress-strain
stances are hydrolyzed and intercellular adhesion is reduced (Greve
curves, as the strain to failure approaches 0.5 or more, the defini-
et al., 1994a; Ng and Waldron, 1997; Lillford, 2001). Thus, cooking
tion of both stress and strain needs to change from its usual basis in
causes fluid to invade the intercellular space and separate the cells
the original dimensions of the object to those at the point of
(Greve et al., 1994b; Thiel and Donald, 2000; Lillford, 2001), which
measurement (Ashby and Jones, 1996).
will make roasted tubers more compliant and lower their elastic
A much more important property for food breakdown in the
modulus. Additionally, relaxed cellecell bonds and lower turgor
mouth (other than the characterization of fracture onset) is that of
pressure cause fractures in most cooked vegetables to run between
the resistance to fractures once they have initiated. The property
cells instead of through them, thereby reducing fracture resistance
that controls the progress of a fracture is toughness, quantified here
(Ng and Waldron, 1997; Lillford, 2001). Therefore, roasting tubers
as the amount of work necessary to grow a crack of a given area.
should reduce both toughness and fracture stress. Additionally,
Since most food objects have inherent flaws and weaknesses prior
cooking gelatinizes starch (Alvarez and Canet, 2001) and in higher
to being processed either orally or manually, toughness is actually
starch tubers this may further decrease fracture resistance if starch
the dominant measure of fracture force resistance in foods (as well
granules lie in the fracture path.
as in a wide range of engineering materials) (Ennos, 2012).
Meat Muscle tissue is approximately 75% water and 20% protein,
A final food parameter of importance is energy dissipation. Stiff
with the majority (~85%) of the protein component made up of
food objects, such as thick and turgid tubers, are likely to fracture in
contractile muscle fibers (myofibrillar and sarcoplasmic proteins).
bending away from cusps, but softer ones will deform and mold
The remaining protein constituent is connective tissue, which
themselves to the working surface of the teeth. In these cases, the
surrounds muscle fibers and hierarchically groups them into lon-
only option for fracture inside the mouth is cuspal penetration.
gitudinal bundles (Tornberg, 2005). Disruption of the perimysium
However, many softer, tougher foods dissipate energy internally
surrounding fiber bundles is probably the most important factor for
making it impossible to ‘run’ a crack ahead of a cusp. Thus, upper
oral processing (Purslow, 1991b). Pounding with a stone tool such
and lower teeth must eventually meet if there is to be a chance for
as an Oldowan hammerstone will mechanically disorganize the
food particle size reduction. The ability of foods to dissipate energy
uniform arrangement of the fiber bundles and may also break the
can be measured as the strain energy lost in viscous/plastic
fibers themselves. Tenderizing in this manner should therefore
behavior, which is the area within the hysteresis loop during
reduce both the fracture stress and toughness of meat.
loading-unloading cycles (Oyen-Tiesma and Cook, 2001). This
Heat denatures (unfolds) and degrades muscle proteins, dam-
measurement is particularly important for evaluating animal soft
ages cell membranes and shrinks (dehydrates) the muscle fibers
tissues such as meat, and may help to predict the probability of
(Lewis and Purslow, 1989; Tornberg, 2005). Tensile strength of meat
fracture in the oral cavity, as less energy dissipation should result in
heated in a water bath increases in two stages, from ~40 to 50  C
158 K.D. Zink et al. / Journal of Human Evolution 77 (2014) 155e166

due to collagen denaturation, and then again from 60 to 90  C as Mechanical processing (tenderization) A replica of an Oldowan
muscle fibers denature and shrinkage and water loss occurs, which hammerstone (a 560.9 g, hand-sized oval rock with smooth edges)
increases protein concentration (Lewis and Purslow, 1989; was used to mechanically tenderize the food. The medulla of each
Christensen et al., 2000; Tornberg, 2005). The morphological and tuber was sliced into 13 mm cubes, then each cube hit six times
chemical changes from heating also increase muscle stiffness with the hammerstone. Goat was cut into a 50 g steak and hit 20
(Lewis and Purslow, 1989; Tornberg, 2005) and fracture strain times. This type of simple tenderization affected tubers and meat
(Willems and Purslow, 1997). Taken together, these effects in boiled differently. The tubers fractured into numerous, relatively large
meat lead to the prediction that dry roasting will increase meat pieces, while the meat remained intact, but was significantly
toughness, fracture stress and strain, and stiffness. Increased ‘mashed’ and the muscle fibers disorganized where struck by the
evaporation from open air cooking may further promote heat- hammerstone.
related water loss and shrinkage, leading to an even stiffer and Roasting Samples were roasted on a warmed-up, tabletop pro-
more fracture resistant material. As meat becomes stiffer, elastic pane grill (Perfect Flame™) with the lid open and the gas flow valve
energy loss is reduced and toughness of the food might decrease as set to ‘high’. Although hominins would not have used grills,
less external work is required for fracture. roasting in this manner is similar to an open fire (the flames are
Mechanical processing This study also tests the general hy- immediately below the grill surface) and allowed cooking of mul-
pothesis that mechanical processing of foods provides sufficient tiple samples at a consistent temperature. Internal temperatures
reduction of masticatory effort to explain the smaller postcanines were measured on selected samples immediately post-cooking
of H. erectus compared with H. habilis and gracile australopiths. (tubers) or during cooking (meat) with a needle probe and digital
Although tooth crown size varies within hominin species, post- thermocouple (Thermoworks™, accuracy ± 0.1  C). Pre- and post-
canine crown area was on average 21e25% smaller in H. erectus roasted weights were recorded (digital scale, accuracy ± 0.1 g).
than in H. habilis and gracile australopiths, respectively, and Homo Before being roasted, each tuber was cut into 17 mm-thick
sapiens molars and premolars were a further 15% smaller (Brace, transverse slices. The slices were then roasted for 15.0 min, 7.5 min
1967; Wolpoff, 1973; McHenry, 1994; McHenry and Coffing, on each side, with each slice rotated to a different spot on the grill
2000). Lucas (2004) modeled postcanine occlusal area reductions every 2.5 min during the cooking process to ensure even heating
in terms of a predicted decrease in food toughness based on frac- despite grill surface temperature variations. This protocol heated
ture scaling. If the fracture of a food is prominently elastic, as food beets to 78.6 ± 2.2  C, carrots to 78.5 ± 1.1  C, and yams to
affected by tenderizing or cooking probably becomes, then Lucas 89.0 ± 2.7  C, and produced a 22 ± 3%, 27 ± 3% and 17 ± 2%
(2004) predicted that dental crown size should scale with the reduction in slice weight (i.e., water loss), respectively.
square root of typical food toughness. Therefore, as processing re- Two pieces of meat, one from the neck muscles (A) and the other
duces the toughness of typical foods in the diet, a concomitant from the knee flexors (B), were roasted on the center of the grill.
decrease in dental occlusal area is predicted to follow, i.e., (tough- They were periodically turned over and were cooked until ‘well-
nessprocessed food/toughnessraw food)0.5 ¼ (dental sizelater hominin/dental done’ and only slightly pink internally (internal temperature,
sizeearlier hominin). This equation leads to the expectation that me- A ¼ 72.2  C, B ¼ 76.0  C). The two steaks differed greatly in size
chanical tenderization caused by pounding will reduce the average (Raw weight: A ¼ 67.7 g, B ¼ 331.7 g) and the larger took almost
toughness of foods 38e44%, allowing for the roughly 21e25% twice as long to reach ‘well-done’ (A ¼ 24 min, B ¼ 46 min). Water
reduction in postcanine occlusal area of H. erectus relative to loss was similar for both steaks (A ¼ 35%, B ¼ 40%).
H. habilis and gracile Australopithecus species. A further reduction in
toughness on top of this (~27%) caused by cooking allowed for the Material property testing
approximately 15% smaller teeth of H. sapiens. If food toughness
decreases with cooking, but not mechanical tenderization, the A Darvell™ HKU portable mechanical property tester (Darvell
argument for early adoption of cooking by H. erectus is supported. et al., 1996) and an Instron™ 5564 were used to measure the ma-
terial properties of the food samples. Details of the particular tester
Methods attachments that were used for each material property test can be
found within the subsections below. Tester attachments were
Food samples cleaned with alcohol between each trial. Unless noted, ten samples
of each raw, mechanically processed and roasted food type were
Organic tubers, red beets (Beta vulgaris), carrots (Daucus carota) tested. Toughness was the only property measured for tenderized
and jewel yams (Ipomoea batatas), were purchased from a local tubers and meat because the processing procedure produced highly
store and stored at 4  C for no more than four days prior to pro- fragmented pieces that precluded additional testing. Each sample
cessing and material property testing. For the meat portion of the was measured with digital or dial (meat tensile tests only) calipers
experiment, an adult goat (Capra aegagrus) was purchased from a (accuracy, ±0.01 and ±0.02 mm, respectively). Raw and roasted
local farm and slaughterhouse (Blood Farms, Groton, MA) and the meat samples were tested both parallel and perpendicular to fiber
fresh carcass transported on ice to the Skeletal Biology Lab, Harvard direction. Because tuber material properties can change drastically
University. Muscle groups were removed using aseptic procedures, depending on the region of the food sampled, testing was only
sealed in vacuum bags and stored at 20  C. Meat was defrosted performed on the inner region (medulla). Samples were wrapped
completely at 4  C for approximately 12e24 h prior to data in damp paper towels to minimize water loss and stored in plastic
collection. bags at 4  C until immediately before testing.
Toughness (tubers) Two experiments were performed, one
Processing procedure testing the effect of roasting and the other testing the effect of
mechanical tenderization. Toughness was measured using a 15
To limit material property variation caused by factors other than included angle wedge fitted onto the Darvell tester (100N load cell).
processing, each tuber was divided into two portions (one was kept The wedge allows for controlled crack growth in semi-brittle foods,
raw, and the other was roasted or tenderized) and the same muscle such as tubers (Vincent et al., 1991).
regions were used for each meat experiment (neck muscle for Experiment #1. Raw and roasted tuber samples were cut into
toughness tests and knee flexors for the tensile tests). rectangular blocks approximately 12e14 mm wide. A wedge was
K.D. Zink et al. / Journal of Human Evolution 77 (2014) 155e166 159

lowered into each sample and stopped when it just pierced the Tensile tests (meat) A bacon slicer was used to cut uniform
material. At this point, the wedge was continuously lowered at a samples of raw and roasted meat. The thickness (t) and width (w) of
rate of ~30 mm/min, while recording the force (F) and resulting each strip (approximately 2 mm and 11 mm, respectively) were
displacement (u) every 0.001 mm, for a final slice depth of recorded. Samples were then placed into an Instron tester (50 N
5.000 mm. The wedge was removed from the sample and rein- load cell) using two pneumatic grips 30 mm apart (lo) and
serted into the previously formed crack to measure the work not increasingly loaded in tension (displacement rate 90 mm/min).
used in crack formation (i.e., work due to friction and elastic Two sets of experiments were performed, a fracture test and a
bending of material against the crack faces). This work was sub- cycling test. Force (F) and displacement (u) were recorded every
tracted from the original work value. After testing, the length of the 0.050 s for the fracture test and 0.034 s for cycling trials.
crack in the horizontal plane (i.e., crack width (w)) was measured, Experiment #1. Each sample was loaded until the meat had
and toughness for each sample was calculated from when the completely fractured into two pieces. As for tubers, the force-
wedge was 2 mm into the sample to when it had progressed to a displacement data were converted into true stress and strain by
depth of 4 mm (i.e., crack depth, m ¼ 2 mm) assuming the incompressibility of muscle in its passive state
(Toughness ¼ (Workinitial pass  Worksecond pass)/wu). (Van Loocke et al., 2006). Fracture stress, fracture strain and the
Experiment #2. Experimental set-up and toughness calculations elastic modulus at 20%, 40%, 60% and 80% fracture stress were
were the same as in experiment #1. The only differences were that recorded. One sample (raw meat, force parallel to fiber direction)
the samples were not a uniform size and data collection began after did not completely fracture and was omitted from analyses
the entire width of the sample was in contact with the wedge (Stress ¼ F(lo þ u)/twlo) (Strain ¼ ln((lo þ u)/lo)).
(samples were the larger of the pieces resulting from the tenderi- Experiment #2. Each sample was cycled to a peak strain of 4%, 8%,
zation process and the tops were irregular). Additionally, because 12% and 16%, and unloaded back to the initial position between
they were smaller than the raw samples, maximum wedge depth each successively large strain loop. Fractures did not initiate during
was 1.5 mm and the toughness for each tenderized piece was these trials. The percent work lost on unloading for each loop was
calculated between a depth of 0.3 and 1.3 mm (i.e., crack depth, calculated. Five samples of each food type (raw and roasted, parallel
u ¼ 1 mm). and perpendicular to muscle fiber direction) were tested.
Toughness (meat) Meat toughness was measured using Swis-
sors™ (Wenger, Switzerland), a pair of precision-manufactured Analyses
tailoring scissors, fitted onto the Darvell tester (100 N load cell).
Cutting tests enable measurements of controlled crack growth in All calculations were performed in Excel (Microsoft 2007),
floppy (high elastic) materials such as meat (Atkins and Mai, 1979; StatView statistical package (SAS Institute, version 5.0.1), and
Lucas et al., 2001). custom written programs in LabView 8.1. In order to avoid as-
Using a bacon slicer, strips of raw and roasted meat of an even sumptions of normal distributions in the data, ManneWhitney U
2e4 mm thickness, and 12e16 mm in width, were cut. Tenderized tests were used to compare the raw and processed food values.
samples were trimmed using a razor blade and gently formed into Significance was set at p  0.05.
similar sized strips. Each sample was placed between the scissor
blades, and the handle depressed at a displacement rate of ~30 mm/ Results
min, while recording the force (F) and resulting displacement (u)
every 0.001 mm until the sample was completely cut. The thickness Toughness (Table 1)
(t) and length (l) of each cut was measured and the work required
to make the cut was calculated. Work due to friction was calculated Roasting reduced the toughness of tubers by 49%, from an
by closing the scissors with no sample between the blades. This average of 1034 ± 345 J/m2 (raw) to 526 ± 120 J/m2 (roasted)
work was then subtracted from the original work value to calculate (p < 0.0001). Mechanically tenderizing the tubers produced a
toughness (Toughness ¼ (Workinitial pass  Worksecond pass)/tl). similar result; toughness decreased 42% from 1080 ± 167 J/m2 (raw)
Compression tests (tubers) A cork borer was used to cut uniform to 622 ± 297 J/m2 (tenderized) (p < 0.0001). Tenderization reduced
cylinders (radius ~3 mm, height ~8 mm) from the center of each beet toughness more than roasting (tenderization ¼ 65%;
tuber, parallel to the long axis of the root. The radius (r) and length roasting ¼ 56%), while the converse was true for yams
(lo) of each sample was recorded. The fracture stress and strain were (roasting ¼ 48%; tenderization ¼ 28%). The percent decrease in
then recorded by continuously compressing each sample between carrot toughness was the same regardless of whether it was roasted
two plates (displacement rate ~32 mm/min) on the Darvell tester (39%) or tenderized (38%).
(1000 N load cell), recording the force (F) e displacement (u) In contrast to the results for tubers, roasting increased the
relationship every 0.001 mm. The data were then converted into toughness of meat by an average of 77% (raw average ¼ 216 ± 78 J/
true stress and strain by assuming conservation of volume (i.e., a m2; roasted average ¼ 381 ± 180 J/m2; p < 0.01). Toughness varied
Poisson's ratio of 0.5, a common and convenient assumption in food greatly depending on whether it was measured parallel or
analysis and supported by measurements such as those by Finney perpendicular to fiber direction. The former primarily measures the
and Hall (1967) on potato). ‘True’ stress and strain were necessary toughness of the weaker connective tissues sheaths surrounding
because as failure strain approaches 0.5 or more, the definition of the fibers, while the latter is a measure of the tougher muscle fibers
both stress and strain needs to change from its usual basis in the themselves, and although roasting increased both of these values
original dimensions of the object to those at the point of mea- (raw parallel ¼ 154 ± 47 J/m2; roasted parallel ¼ 218 ± 49 J/m2;
surement (Ashby and Jones, 1996). Strain was then converted to p ¼ 0.01) (raw perpendicular ¼ 277 ± 46 J/m2; roasted
absolute measures; a negative value obtained in the calculation perpendicular ¼ 545 ± 84 J/m2; p < 0.001), its effect was greatest
(below) signifies a compressive state. Peak stress (fracture stress) perpendicular to fiber direction (perpendicular ¼ 97% increase;
and the corresponding strain were recorded, and the elastic parallel ¼ 41% increase).
modulus (slope of the line) was calculated at 20%, 40%, 60% and 80% Unlike with tubers, tenderizing did not affect meat toughness.
fracture stress. Multiple measures are needed because a linear Although tenderized meat (163 ± 68 J/m2) was less tough than raw
stressestrain relationship is not anticipated for all foods meat measured across the muscle fibers (p < 0.01), it did not differ
(Stress ¼ F(lou)/pr2lo) (Strain ¼ ln((lou)/lo)). significantly from the toughness of raw meat measured parallel to
160 K.D. Zink et al. / Journal of Human Evolution 77 (2014) 155e166

Table 1 greatest for yams (69%), followed by carrots (20%), and beets (11%).
Average toughness (J/m2) of raw, roasted and tenderized tubers and goat meat. Conversely, the average compressive strain required for fracture
Treatment Toughness (J/m2)a increased by 58%, going from 0.33 ± 0.09 in raw tubers to
Tubers, Experiment #1
0.52 ± 0.19 with roasting (p < 0.0001). As with fracture stress, the
Raw beet (n ¼ 10) 1424.9 (208.8) greatest effect was for yams (74%), although beets followed closely
Roasted beet (n ¼ 10) 622.6 (120.0) (70%). Carrot fracture strain, in comparison, increased a modest 34%
Raw carrot (n ¼ 10) 903.8 (256.0) with roasting.
Roasted carrot (n ¼ 10) 553.0 (48.1)
Inspection of the stress-strain curves for raw and roasted tubers
Raw yam (n ¼ 10) 774.6 (101.3)
Roasted yam (n ¼ 10) 403.0 (39.4) showed that the curves changed in shape from linear/slightly r-
All tubers, raw (n ¼ 30) 1034.4 (344.7) shaped (raw) to J- or even slightly s-shaped (roasted) (Fig. 2).
All tubers, roasted (n ¼ 30) 526.2 (119.8) Comparing modulus values, roasting decreased the stiffness of all
Tubers, Experiment #2
tubers (45% reduction, tuber and modulus average; p < 0.0001),
Raw beet (n ¼ 10) 922.7 (103.5)
Tenderized beet (n ¼ 10) 322.2 (90.5)
with the exception of carrots measured at 40% fracture stress (no
Raw carrot (n ¼ 10) 1239.7 (134.0) significant difference). Average raw tuber modulus started off high
Tenderized carrot (n ¼ 10) 769.1 (317.1) at 6026 ± 2087 kPa (measured at 20% fracture stress) and declined
Raw yam (n ¼ 10) 1078.5 (71.3) steadily to 4627 ± 1095 kPa (measured at 80% fracture stress). In
Tenderized yam (n ¼ 10) 773.6 (162.3)
contrast, roasted tubers followed a different pattern, starting off at a
All tubers, raw (n ¼ 30) 1080.3 (166.8)
All tubers, tenderized (n ¼ 30) 621.6 (297.1) low of 2449 ± 1332 kPa and then increasing to 3174 ± 1582 kPa,
Meatb before decreasing to 3024 ± 1535 kPa and 2629 ± 1491 kPa
Raw, parallel (n ¼ 10) 154.3 (46.5) (measured at 20%, 40%, 60% and 80% fracture stress, respectively).
Roasted, parallel (n ¼ 10) 218.2 (48.7) Yams had the greatest reduction in overall stiffness with roasting
Raw, perpendicular (n ¼ 10) 277.1 (46.3)
Roasted, perpendicular (n ¼ 10) 544.6 (84.2)
(76%), followed by carrots (31%) and beets (30%).
Tenderized (n ¼ 10)c 163.3 (68.0)
a
Toughness of tubers and meat was measured using wedge and cutting tests,
respectively. One standard deviation is in parentheses. Underlined and bolded Meat tensile tests (Table 2)
values indicate a significant difference between processed foods and their raw
counterparts (p  0.05, ManneWhitney U test). See text for food processing details. Unlike tubers, roasting increased the fracture stress of meat,
b
Meat was measured both parallel and perpendicular to muscle fiber direction. with the greatest change occurring parallel to fiber direction.
c
Toughness of tenderized meat differed significantly from raw meat measured
perpendicular, but not parallel to muscle fiber direction.
Maximum stress at fracture went from 20 ± 10 kPa (raw) to
31 ± 9 kPa (roasted), when meat was pulled perpendicular to the
direction of the fibers, primarily breaking the weaker connective
fiber direction or average raw meat toughness (average raw meat tissue rather than the fibers themselves (p ¼ 0.03). When meat was
toughness parallel and perpendicular to fiber direction, 216 ± 78 J/m2). tensed parallel to fiber direction, fracture stress rose from
96 ± 32 kPa (raw) to 309 ± 117 kPa (roasted), a 222% increase
(p < 0.001). The maximum amount of tensile deformation required
Tuber compression tests (Table 2) to fracture the meat into two separate pieces did not significantly
change with roasting (raw parallel ¼ 0.65 ± 0.16; roasted
With one exception (noted below), roasting significantly parallel ¼ 0.48 ± 0.20) (raw perpendicular ¼ 0.54 ± 0.11; roasted
affected the fracture stress, fracture strain and stiffness of tubers. In perpendicular ¼ 0.49 ± 0.12).
all cases, roasting altered yams more than carrots and beets. Roasting changed the shape of the stress-strain curve from
The stress required to fracture tubers was greatly lowered with linear or slightly J-shaped to r-shaped, but only when pulling par-
roasting, decreasing 28% from a raw average of 1349 ± 349 kPa to a allel to fiber direction (Fig. 3). Curve shape remained the same (J-
roasted average of 974 ± 555 kPa (p < 0.01). This reduction was shaped) when meat was pulled perpendicular to fiber direction,

Table 2
Average fracture stress (kPa) and strain, and modulus of elasticity (kPa) of raw and roasted tubers and goat meat.a

Treatment Fracture stress (kPa) Fracture strain Modulus at 20% Modulus at 40% Modulus at 60% Modulus at 80%
fracture stress (kPa)b fracture stress (kPa)b fracture stress (kPa)b fracture stress (kPa)b

Tubers
Raw beet (n ¼ 10) 1610.1 (130.8) 0.43 (0.05) 4579.0 (1076.3) 4090.3 (439.2) 4159.6 (431.3) 4661.9 (692.0)
Roasted beet (n ¼ 10) 1427.7 (201.0) 0.73 (0.06) 2737.8 (461.3) 3385.3 (646.0) 3311.6 (690.0) 2843.2 (1028.1)
Raw carrot (n ¼ 10) 1531.1 (199.7) 0.32 (0.03) 6816.4 (2520.0) 5982.8 (1328.4) 5659.7 (907.3) 5570.5 (866.2)
Roasted carrot (n ¼ 10) 1217.8 (299.1) 0.43 (0.07) 3527.7 (1326.8) 4674.0 (1316.7) 4498.5 (976.7) 3943.0 (1019.1)
Raw yam (n ¼ 10) 905.8 (59.0) 0.23 (0.04) 6682.3 (1740.2) 5583.7 (1237.2) 4411.7 (785.0) 3649.8 (764.9)
Roasted yam (n ¼ 10) 277.1 (166.2) 0.40 (0.18) 1081.2 (525.6) 1463.9 (337.4) 1262.9 (455.9) 1102.0 (713.6)
All tubers, raw (n ¼ 30) 1349.0 (348.5) 0.33 (0.09) 6025.9 (2087.1) 5218.9 (1330.0) 4743.7 (974.4) 4627.4 (1095.3)
All tubers, roasted (n ¼ 30) 974.2 (554.8) 0.52 (0.19) 2448.9 (1331.6) 3174.4 (1581.9) 3024.4 (1534.9) 2629.4 (1491.2)
Meatc
Raw, parallel (n ¼ 9)d 96.0 (32.4) 0.65 (0.16) 148.6 (65.4) 148.3 (47.4) 246.5 (59.7) 328.8 (106.7)
Roasted, parallel (n ¼ 10) 309.2 (116.6) 0.48 (0.20) 1066.7 (580.1) 935.7 (556.2) 822.4 (369.6) 733.6 (166.9)
Raw, perpendicular (n ¼ 10) 20.3 (9.8) 0.54 (0.11) 53.1 (41.3) 66.1 (50.0) 65.7 (52.7) 78.6 (45.6)
Roasted, perpendicular (n ¼ 10) 30.7 (8.9) 0.49 (0.12) 55.8 (20.6) 90.6 (25.3) 84.6 (40.5) 111.8 (39.4)
a
Properties of tubers and meat were measured using compression and tensile tests, respectively. One standard deviation is in parentheses. Underlined and bolded values
indicate a significant difference between raw and roasted foods (p  0.05, ManneWhitney U test). See text for food processing details.
b
Elastic modulus was measured at 20%, 40%, 60% and 80% fracture stress.
c
Meat was measured both parallel and perpendicular to muscle fiber direction.
d
One raw meat sample (measured parallel to fiber direction) did not completely fracture and was omitted from the analyses, reducing sample size to nine.
K.D. Zink et al. / Journal of Human Evolution 77 (2014) 155e166 161

A A
2000 250
Beet Parallel
Parallel
1600 200
Stress (kPa)

1200

Stress (kPa)
150
800
100
400
50
0
0 0.2 0.4 0.6 0.8 1
0
Strain
0 0.1 0.2 0.3 0.4 0.5 0.6
B
2000 Strain
Carrot
1600 B
25
Stress (kPa)

Perpendicular
1200
20
800

Stress (kPa)
15
400
10
0
0 0.2 0.4 0.6 0.8 1 5
Strain
C 0
2000
Yam 0 0.1 0.2 0.3 0.4 0.5 0.6
1600 Strain
Stress (kPa)

1200 Figure 3. Average stress-strain curves of raw (light) and roasted (dark) meat. Meat was
loaded in tension both parallel (A) and perpendicular (B) to muscle fiber direction.
800

400
samples tensed to 4%, 8%, 12% and 16% strain, respectively). There
0 was no significant difference in energy loss between raw and
0 0.2 0.4 0.6 0.8 1 roasted samples pulled perpendicular to fiber direction. Although
Strain roasting had no effect, average energy loss measured in this di-
rection was relatively low and comparable to that of roasted meat
Figure 2. Representative stressestrain curves of raw (light) and roasted (dark) beets
measured parallel to fiber direction (45 ± 7% versus 45 ± 6%,
(A), carrots (B) and yams (C). Fracture occurred at the last point displayed.
respectively).

regardless of whether it was raw or roasted. A similar pattern Discussion


emerges for comparisons of stiffness. The average modulus of
roasted meat parallel to fiber direction was 308% higher than raw As predicted, tuber material properties responded to both me-
meat (p < 0.001), while the modulus measured perpendicular to chanical tenderization and roasting. Tenderizing reduced tuber
fiber direction did not significantly change. Parallel to fiber direc- toughness by 42%, and roasting reduced toughness slightly more,
tion, the raw meat modulus was unchanged at 20% and 40% fracture 49% on average. Roasting the tubers also decreased fracture stress,
stress (149 ± 65 kPa and 148 ± 47 kPa, respectively), and then made them more compliant, and increased fracture strain. In
increased to 247 ± 60 kPa and 329 ± 107 kPa (measured at 60% and contrast, meat did not respond as predicted for all material prop-
80% fracture stress, respectively). The comparable roasted meat erty measures. Tenderizing the meat by pounding with a ham-
modulus, however, steadily decreased with increasing fracture merstone did not change toughness, suggesting that this form of
stress, going from a high of 1067 ± 580 kPa to 936 ± 556 kPa, processing simply disorganized the muscle fibers and connective
822 ± 370 kPa and 734 ± 167 kPa (measured at 20%, 40%, 60% and tissue without altering their intrinsic properties. Additionally,
80% fracture stress, respectively). Regardless of whether the meat although roasting significantly increased meat toughness, fracture
was raw or roasted, stiffness perpendicular to the direction of the stress and stiffness, and reduced elastic energy loss, the latter two
fibers was much lower than when measured parallel to fiber di- properties responded only when measured parallel to fiber direc-
rection (Raw ¼ perpendicular 70% lower than parallel; tion. Finally, contrary to predictions, roasting meat did not signifi-
Roasted ¼ perpendicular 90% lower than parallel; p < 0.001). cantly affect the amount of strain necessary for fracture.
What are the implications of these material property data for
Meat cycling (Table 3) the evolution of hominin mastication? The tuber results support
the hypothesis that mechanical processing could have played a role
Roasting significantly decreased work loss (energy dissipation), in favoring the smaller postcanine crowns of H. erectus compared
but only when measured parallel to fiber direction (p < 0.01) with H. habilis and gracile australopiths. Based on fracture scaling,
(Fig. 4). On average, roasted samples lost 28% less energy than raw which predicts that dental size is proportional to the square root of
samples (66 ± 11% versus 38 ± 3%, 74 ± 11% versus 44 ± 4%, 77 ± 10% typical food toughness (Lucas, 2004), the 42% average toughness
versus 49 ± 3%, 78 ± 9% versus 51 ± 3% for raw versus roasted decrease measured in tubers pounded with a hammerstone should
162 K.D. Zink et al. / Journal of Human Evolution 77 (2014) 155e166

Table 3
Average percent work lost (energy dissipation) during loading-unloading cycles of raw and roasted meat.a

Parallel (n ¼ 5) Perpendicular (n ¼ 5)

Raw Roasted Raw Roasted


5 5 4 4 5 6
4% Strain Work (J), loaded 7.03  10 (2.23  10 ) 2.34  10 (1.21  10 ) 1.50  10 (9.99  10 ) 3.27  105 (1.26  105)
Work lost (J), unloaded 4.62  105 (1.77  105) 8.87  105 (4.64  105) 6.21  106 (3.49  106) 1.44  105 (6.68  106)
% Work Lostb 65.6% (11.3%) 37.9% (2.5%) 43.6% (6.8%) 43.4% (7.6%)
8% Strain Work (J), loaded 2.46  104 (6.59  105) 1.02  103 (4.93  104) 6.42  105 (4.73  105) 1.11  104 (3.76  105)
Work lost (J), unloaded 1.79  104 (4.73  105) 4.57  104 (2.26  104) 2.60  105 (2.41  105) 5.00  105 (2.35  105)
% Work lostb 73.6% (11.0%) 44.3% (4.4%) 37.9% (7.4%) 43.8% (6.8%)
12% Strain Work (J), loaded 4.64  104 (1.34  104) 2.33  103 (1.06  103) 1.53  104 (1.05  104) 2.23  104 (6.17  105)
Work lost (J), unloaded 3.48  104 (7.41  105) 1.14  103 (5.25  104) 6.37  105 (5.97  105) 1.05  104 (4.65  105)
% Work Lostb 76.9% (9.9%) 48.8% (3.4%) 37.7% (10.8%) 45.9% (8.2%)
16% Strain Work (J), loaded 6.62  104 (2.18  104) 3.90  103 (1.69  103) 2.87  104 (1.78  104) 3.55  104 (6.99  105)
Work lost (J), unloaded 5.01  104 (1.14  104) 1.99  103 (8.52  104) 1.12  104 (8.36  105) 1.73  104 (6.17  105)
% Work lostb 77.8% (8.9%) 50.9% (3.0%) 36.8% (8.6%) 47.8% (8.0%)
a
Meat was measured both parallel and perpendicular to muscle fiber direction. Each sample was successively loaded-unloaded to 4%, 8%, 12% and 16% strain. One standard
deviation in parentheses. See text for food processing details.
b
Percent work loss was calculated as the difference between the work performed during loading and unloading of the sample (i.e., work lost) divided by the work per-
formed during loading. Underlined and bolded values indicated a significant difference in percent work lost between raw and roasted meat (p  0.05, ManneWhitney U test).

result in a 24% decrease in postcanine occlusal area, a reduction that higher starch content and the presence of a unique heat-activated
corresponds remarkably closely with observations from the fossil beamylase that further aids starch gelatinization (Binner et al.,
record (e.g., Brace, 1967; Wolpoff, 1973; McHenry, 1994; McHenry 2000). In contrast to yams, beet toughness was more affected by
and Coffing, 2000). Because pounding breaks tubers into smaller tenderization than roasting (65% and 56% decrease, respectively),
pieces, it also reduces ingested particle size, which may have and carrot toughness was equally affected by the two processing
further decreased required masticatory force. methods (38% and 39% decrease, respectively). Data on additional
Roasting tubers decreased toughness 17% more than tenderi- foods are needed before we can conclude with confidence that the
zation. Although the average toughness difference caused by the effects of cooking on tuber toughness (and masticatory force) are
two processing methods was not as large as predicted from fracture superior to that of tenderization.
scaling (27%), the larger effect of roasting compared with tenderi- Agrawal et al. (1998) demonstrated a negative correlation be-
zation suggests that habitual cooking may have contributed to the tween chewing muscle recruitment and a food's (toughness/stiff-
dental size reductions and facial gracility of later Homo. It is ness)0.5. Thus, across a wide-range of food types, foods with lower
important to note, however, that while roasting generally reduced stiffness relative to toughness are chewed with less muscular effort.
the toughness of tubers more than tenderization, this result was To further explore the differential effects of cooking, Table 4 presents
driven largely by the yam data; the effects of roasting on tuber calculations of this index for the three tubers (N.B. this index could
material properties were largest for yams, possibly due to a higher not be calculated for tenderized tubers because we did not measure
internal temperature reached during cooking, and also from their the stiffness of these foods.). While (toughness/stiffness)0.5

Figure 4. Left: Average force-displacement curves for raw (light) and roasted (dark) meat. Each sample was successively loaded-unloaded to 4%, 8%, 12% and 16% strain. Right:
Percent work lost (energy dissipation) during the unloading phase of each work cycle. Meat was loaded in tension both parallel (A) and perpendicular (B) to muscle fiber direction.
Box plot whiskers extend to the 10th and 90th percentiles. *p  0.05, ManneWhitney U Test.
K.D. Zink et al. / Journal of Human Evolution 77 (2014) 155e166 163

Table 4 fracture at all, but was instead simply mashed. Cooking, however,
Ratio of the toughness and stiffness (elastic modulus) of raw and roasted tubers and may improve masticatory performance by increasing masticatory
meat.a
efficiency. We calculated (toughness/stiffness)0.5 for the material
Treatment (Toughness/Stiffness)0.5 properties that primarily describe the meat connective tissue
Tubers (toughness and average stiffness measured parallel and perpen-
Raw beet (n ¼ 10) 0.57 dicular to fiber direction, respectively) and the meat muscle fibers
Roasted beet (n ¼ 10) 0.45 (toughness and average stiffness measured perpendicular and
Raw carrot (n ¼ 10) 0.39
parallel to fiber direction, respectively) (Table 4). While the
Roasted carrot 0.36
Raw yam (n ¼ 10) 0.39 (toughness/stiffness)0.5 of meat connective tissue changed only
Roasted yam (n ¼ 10) 0.57 slightly with roasting (~4% increase), it decreased nearly a third
Meatb (~31%) for the muscle fibers. Although a decrease in (toughness/
Raw, connective tissue (n ¼ 10) 1.53
stiffness)0.5 predicts an increase in masticatory force per chew,
Roasted, connective tissue (n ¼ 10) 1.60
Raw, muscle fiber (n ¼ 9)c 1.13
these increases may be mitigated by an increase in the rate at
Roasted, muscle fiber (n ¼ 10) 0.78 which fractures are produced by the teeth (Agrawal et al., 1997,
a 1998). Improved comminution is likely to lead to a significant
See text for experimental details. Stiffness (elastic modulus) at 20%, 40%, 60%
and 80% fracture stress was averaged. reduction in the number of chews required, and therefore lower
b
Material properties of meat were measured both parallel and perpendicular to overall masticatory effort (i.e., masticatory force per
muscle fiber direction. (Toughness/Stiffness)0.5was calculated for the properties that chew  number of chews).
primarily describe 1) the connective tissue of the meat (i.e., toughness measured
Another way in which cooking may have helped hominins
parallel to muscle fiber direction, and average stiffness measured perpendicular to
muscle fiber direction), and 2) the muscle fibers of the meat (i.e., toughness
chew meat is that roasting stiffens meat parallel to fiber direc-
measured perpendicular to muscle fiber direction, and average stiffness measured tion and decreases energy loss (hysteresis). Although fractures in
parallel to muscle fiber direction). meat tend to run preferentially through the weaker perimysial
c
One raw meat sample (measured parallel to muscle fiber direction) did not (connective tissue) bonds as opposed to breaking the tougher
completely fracture during the tensile tests and therefore stiffness could not be quan-
muscle fibers (Purslow, 1985; Lewis and Purslow, 1990; Willems
tified. This sample was omitted from the analyses, and sample size was reduced to nine.
and Purslow, 1997; Lillford, 2001), when muscle fiber fractures
do develop, comminution efficiency (food breakdown per chew)
increased approximately 47% in roasted yams, it decreased in roas- will increase. The maximum number of particles possible from
ted beets and carrots. This difference leads to the prediction that connective tissue fracturing is positively correlated with the
force per chew will actually be higher in roasted versus raw beets number of muscle fibers present. In a given volume of food, fiber
and carrots. Such force increases may be mitigated, however, by number is greatest in slices running against the grain of the meat
fewer chews overall because increased fragmentation occurs per and lowest in cuts following the longitudinal axis of the muscle.
chew at lower (toughness/stiffness)0.5 (Agrawal et al., 1997). In other In the latter case especially, stiffening with roasting may make
words, foods with higher stiffness and relatively low toughness the food less ‘displacement limited’ and reduce fracture strain.
undergo less deformation prior to fracture. Therefore, in the space- Although ultimate fracture strain was not different between raw
limited environment between occluding teeth, these foods will and roasted meat, there was a clear dip in stress at approxi-
fracture more per chew. Increased fracture rates also explain why mately 30% strain for the roasted meat pulled parallel to fiber
foods of low (toughness/stiffness)0.5 are consumed with more direction (Fig. 3). This dip did not occur in any of the other
muscular effort per chew; as the teeth come into contact with the curves and may indicate fracture initiation, breakage of the first
food they require more force to propagate a larger number of frac- muscle fiber, after which the crack deviates into the weaker
tures within the material. perimysium and eventually moves across the remaining fibers
In addition to material property changes, cooking may also in- (increasing stress) to completely fracture the meat into two
crease the nutritional density of tubers and other underground pieces. A decrease in energy loss with roasting will further aid
storage organs through dehydration, increasing digestibility, and fracture propagation and significantly increase the probability of
lowering cost of digestion (Carmody et al., 2011), all of which might fracture in the limited strain environment between two
further reduce masticatory effort per calorie of food. While we occluding teeth (Fig. 4, Table 3).
focused only on the inner, edible portion of the tubers, Dominy et al. Finally, as with tubers, roasting may make the meat more
(2008) found that cooking also makes removal of the tough, outer nutritionally dense, allowing for less consumption per calorie.
casing of tubers much easier. Improving the ability for hominins to Water loss was significant with roasting, 35e40% by mass, greatly
access the edible portion of underground storage organs would increasing the amount of nutrients per gram. Digestive costs may
have allowed for overall increases in net energy gained from the also improve with cooking as shown in studies of Burmese pythons
food as less effort is required for ingestion. (Boback et al., 2007) and mice (Carmody et al., 2011) fed raw versus
Although the tuber data support the hypothesis that mechani- cooked meat. Cooking also has the added benefit of reducing
cal and thermal processing techniques could have allowed for the endogenous parasites and killing bacteria that are especially
postcanine size decreases within Homo, the meat results are more prevalent in meat (particularly scavenged meat; Ragir et al., 2000)
difficult to interpret. Contrary to predictions, mechanical tenderi- and which might cause illness. Reducing immune and digestive
zation did not affect meat toughness and roasted meat was actually costs will increase the net energy gained from roasted meat
tougher and required approximately 50e220% more stress to (Carmody and Wrangham, 2009), and therefore less food would
fracture than raw meat. Toughness changes, however, do not need to be ingested to attain a certain number of calories.
necessarily mean that more masticatory effort must be expended In conclusion, the results of this study indicate a substantial
to chew raw versus roasted meat. As previously discussed, primate reduction in masticatory effort resulting from the use of simple
molars are poorly adapted to chewing raw meat. Compression does food processing techniques that were available to hominins, espe-
not produce fractures in raw meat, which instead must be sheared cially after the start of the Lower Paleolithic. Tubers and meat
or pulled (this is why cutting and tensile tests were used to mea- responded differently to processing, however, which cautions
sure meat properties.). For example, the meat used for this study against grouping these (and other) foods together when discussing
was hit 20 times with a replica Oldowan hammer stone and did not the potential ramifications of food processing behavior for
164 K.D. Zink et al. / Journal of Human Evolution 77 (2014) 155e166

mastication. It should also be noted that masticatory forces are example, although boiling greatly decreases potato and white
distributed across the occlusal surface of the tooth, and that dental onion toughness, it has no effect on the toughness of white turnips
stress (force per area) is one of the main contributors to food (Lucas, 2004). A number of factors such as the amount of starch
fracture. Eng et al. (2013) calculated maximum bite stress at the and fiber in plants, and age, fat reserves, amount and distribution
second molar in apes and hominins, and found that although dental of connective tissue in muscle tissue will alter fracture properties.
stress remained fairly constant across most taxa (r2 ¼ 0.93), Homo Only four domesticated foods (three tubers and goat meat) were
species fell well below the regression line, largely caused by a tested and although they are representative of food sources
substantial decrease in the amount of maximum bite force that potentially important to early hominins, testing of additional
they could produce. This large residual suggests a grade shift with foods, especially wild vegetation and game meat will further
the genus Homo in which the most mechanically demanding foods inform hominin dietary hypotheses.
that were consumed no longer required high dental stresses,
probably because of the use of food processing techniques to soften Acknowledgments
the foods prior to consumption.
Regardless of the type of processing methods employed, it is We are very grateful to Rachel Carmody, Andrew Cunningham,
reasonable to predict that hominins would choose to processes Carolyn Eng, Tanya Smith and Richard Wrangham for helpful dis-
the foods that were the most mechanically demanding and cussion and comments. We would also like to thank two anony-
required the highest masticatory forces to consume. Therefore, mous reviewers who helped to improve this manuscript. This work
by processing foods, selection would no longer act to maintain was funded by the National Science Foundation through a doctoral
features necessary to generate high occlusal stresses or to dissertation improvement grant (#0925688) awarded to K. Zink,
maintain energetically costly masticatory structures such as large and also by the American School of Prehistoric Research (Peabody
muscles and heavy, robust facial bones (Brace, 1967, 1995; Museum, Harvard University).
Wolpoff, 1971; Wrangham et al., 1999; Teaford et al., 2002;
Klein, 2009; Lieberman, 2011). This leaves open the possibility
for smaller postcanine teeth to evolve for other reasons, such as References
to fit into shorter, less prognathic faces (for review, see Agrawal, K.R., Lucas, P.W., Prinz, J.F., Bruce, I.C., 1997. Mechanical properties of foods
Lieberman, 2011). responsible for resisting food breakdown in the human mouth. Arch. Oral Biol.
42, 1e9.
Agrawal, K.R., Lucas, P.W., Bruce, I.C., Prinz, J.F., 1998. Food properties that influence
Future research
neuromuscular activity during human mastication. J. Dent. Res. 77, 1931e1938.
Alvarez, M.D., Canet, W., 2001. Kinetics of thermal softening of potato tissue heated
The data presented in this study highlight the need to test the by different methods. Eur. Food Res. Technol. 212, 454e464.
effects of food processing on masticatory performance variables. Ashby, M.F., Jones, D.R.H., 1996. Engineering Materials 1: an Introduction to their
Properties and Applications, Second edition. Butterworth-Heinemann, Boston.
While material property tests are both useful and relatively inex- Atkins, A.G., Mai, Y.W., 1979. Guillotining of materials. J. Mat. Sci. 14, 2747e2754.
pensive to perform on a wide variety of foods and processing types, Baird, L., Rodwell, J., 2005. The Complete Book of Raw Food: Healthy, Delicious
the relationship between food properties and fracture in the oral Vegetarian Cuisine Made with Living Foods. Healthy Living Books/Hatherleigh
Press, New York.
cavity is complex. For example, the low toughness and high Beleia, A., Yamashita, F., de Moraes, S.R., da Silveira, C.A., Miranda, L.A., 2004a.
(toughness/stiffness)0.5 of raw meat compared with both raw and Textural changes during cooking of cassava (Manihot esculenta Crantz) roots.
roasted tubers leads to the prediction of less force per chew when J. Sci. Food Agr. 84, 1975e1978.
Beleia, A., Prudencio-Ferreira, S.H., Yamashita, F., Sakamoto, T.M., Ito, L., 2004b.
consuming meat. Therefore, it is reasonable to hypothesize that the Sensory and instrumental texture analysis of cassava (Manihot esculenta Crantz)
addition of more raw meat to hominin diets could have made roots. J. Texture Stud. 35, 542e553.
possible the evolution of smaller, less robust masticatory Ben-Dor, M., Gopher, A., Hershkovitz, I., Barkai, R., 2011. Man the fat hunter: the
demise of Homo erectus and the emergence of a new hominin lineage in the
morphology in the genus Homo by relaxing selection on having
Middle Pleistocene (ca. 400 kyr) Levant. PLoS One 6.
large, thickly-enameled teeth, as well as many features of the skull Berna, F., Goldberg, P., Horwitz, L.K., Brink, J., Holt, S., Bamford, M., Chazan, M., 2012.
that generate and withstand large masticatory forces. A caveat to Microstratigraphic evidence of in situ fire in the Acheulean strata of Wonder-
werk Cave, Northern Cape province, South Africa. Proc. Natl. Acad. Sci. 109,
the material property data, however, is that the specific testing
E1215eE1220.
environments used were conducive to fracture in meat (cutting and Binner, S., Jardine, W.G., Renard, C.M.C.G., Jarvis, M.C., 2000. Cell wall modifications
tensing). As discussed previously, the ability of hominins to during cooking of potatoes and sweet potatoes. J. Sci. Food Agr. 80, 216e218.
comminute this food source efficiently in the raw state is ques- Boback, S.M., Cox, C.L., Ott, B.D., Carmody, R., Wrangham, R.W., Secor, S.M., 2007.
Cooking and grinding reduces the cost of meat digestion. Comp. Biochem.
tionable. As with chimpanzees, considerable chewing time and Physiol. A 148, 651e656.
effort may have been required to consume raw meat, and even Boesch, C., Boesch-Achermann, H., 2000. The Chimpanzees of the Taï Forest:
though the force per chew may have been low, many repetitive Behavioural Ecology and Evolution. Oxford University Press, New York.
Brace, C.L., 1967. Environment, tooth form, and size in the Pleistocene. J. Dent. Res.
chewing cycles would have been necessary (a hypothesis that re- 46, 809e816.
quires further testing). Brace, C.L., 1995. Biocultural interaction and the mechanism of mosaic evolution in
Future research should also focus on testing the material the emergence of modern morphology. Am. Anthropol. 97, 711e721.
Bunn, H.T., 1981. Archaeological evidence for meat-eating by Plio-Pleistocene
properties of other food and processing types relevant to early hominids from Koobi Fora and Olduvai Gorge. Nature 291, 574e577.
hominins. Pounding and dry roasting over an open flame are only Bunn, H.T., 1994. Early Pleistocene hominid foraging strategies along the ancestral
two of many processing types that would have been available to Omo river at Koobi-Fora, Kenya. J. Hum. Evol. 27, 247e266.
Bunn, H.T., 2007. Meat made us human. In: Ungar, P. (Ed.), Evolution of the Human
early hominins, and roasting was performed in this experiment Diet: the Known, the Unknown, and the Unknowable. Oxford University Press,
only to a single, relatively high, temperature. Many other methods New York, pp. 191e211.
such as below-ground slow roasting, soaking, leaching, chemical Bunn, H.T., Kroll, E.M., 1986. Systematic butchery by Plio-Pleistocene hominids at
Olduvai-Gorge, Tanzania. Curr. Anthropol. 27, 431e452.
tenderization and especially slicing, grinding, and slow drying
Carmody, R.N., Wrangham, R.W., 2009. The energetic significance of cooking.
over a small fire or in the hot sun may have been employed as J. Hum. Evol. 57, 379e391.
well. It is also important to note that these methods are not Carmody, R.N., Weintraub, G.S., Wrangham, R.W., 2011. Energetic consequences of
mutually exclusive and multiple processing steps may have been thermal and nonthermal food processing. Proc. Natl. Acad. Sci. 108,
19199e19203.
performed on foods. Additionally, processing techniques have Chamberlain, A.T., Wood, B.A., 1985. A reappraisal of variation in hominid
different effects on the material properties of different foods. For mandibular corpus dimensions. Am. J. Phys. Anthropol. 66, 399e405.
K.D. Zink et al. / Journal of Human Evolution 77 (2014) 155e166 165

Chang, H.J., Wang, Q., Zhou, G.H., Xu, X.L., Li, C.B., 2010. Influence of weak organic Lewis, G.J., Purslow, P.P., 1990. Connective tissue differences in the strength of
acids and sodium chloride marination on characteristics of connective tissue cooked meat across the muscle fibre direction due to test specimen size. Meat
collagen and textural properties of beef semitendinosus muscle. J. Texture Stud. Sci. 28, 183e194.
41, 279e301. Lieberman, D., 2011. The Evolution of the Human Head. Belknap Press of Harvard
Christensen, M., Purslow, P.P., Larsen, L.M., 2000. The effect of cooking temperature University Press, Massachusetts.
on mechanical properties of whole meat, single muscle fibres and perimysial Lillford, P.J., 2001. Mechanisms of fracture in foods. J. Texture Stud. 32, 397e417.
connective tissue. Meat Sci. 55, 301e307. Loucks, L.J., Ray, E.E., Berry, B.W., Leighton, E.A., Gray, D.G., 1984. Effects of me-
Christensen, M., Young, R.D., Lawson, M.A., Larsen, L.M., Purslow, P.P., 2003. Effect of chanical tenderization and cooking treatments upon product attributes of pre-
added m-calpain and post-mortem storage on the mechanical properties of rigor and post-rigor beef roasts. J. Anim. Sci. 58, 626e630.
bovine single muscle fibres extended to fracture. Meat Sci. 66, 105e112. Lucas, P.W., 2004. Dental Functional Morphology: How Teeth Work. Cambridge
Christensen, M., Larsen, L.M., Ertbjerg, P., Purslow, P.P., 2004. Effect of proteolytic University Press, New York.
enzyme activity and heating on the mechanical properties of bovine single Lucas, P.W., Beta, T., Darvell, B.W., Dominy, N.J., Essackjee, H.C., Lee, P.K.D., Osorio, D.,
muscle fibres. Meat Sci. 66, 361e369. Ramsden, L., Yamashita, N., Yuen, T.D.B., 2001. Field kit to characterize physical,
Combes, S., Lepetit, J., Darche, B., Lebas, F., 2003. Effect of cooking temperature and chemical and spatial aspects of potential primate foods. Folia Primatol. 72,
cooking time on Warner-Bratzler tenderness measurement and collagen con- 11e25.
tent in rabbit meat. Meat Sci. 66, 91e96. McHenry, H.M., 1994. Tempo and mode in human evolution. Proc. Natl. Acad. Sci. 91,
Dan, H., Watanabe, H., Dan, I., Kohyama, K., 2003. Effects of textural changes in 6780e6786.
cooked apples on the human bite, and instrumental changes. J. Texture Stud. 34, McHenry, H.M., Coffing, K., 2000. Australopithecus to Homo: transformations in body
499e514. and mind. A. Rev. Anthropol. 29, 125e146.
Darvell, B.W., Lee, P.K.D., Yuen, T.D.B., Lucas, P.W., 1996. A portable fracture McPherron, S.P., Alemseged, Z., Marean, C.W., Wynn, J.G., Reed, D., Geraads, D.,
toughness tester for biological materials. Meas. Sci. Technol. 7, 954e962. Bobe, R., Bearat, H.A., 2010. Evidence for stone-tool-assisted consumption of
Dilek, M., Polat, H., Kezer, F., Korcan, E., 2011. Application of locust bean gum edible animal tissues before 3.39 million years ago at Dikika, Ethiopia. Nature 466,
coating to extend shelf life of sausages and garlic-flavored sausage. J. Food 857e860.
Process. Pres. 35, 410e416. Milton, K., 1999. A hypothesis to explain the role of meat-eating in human evolu-
Dixon, C.L., Woerner, D.R., Tokach, R.J., Chapman, P.L., Engle, T.E., Tatum, J.D., tion. Evol. Anthropol. 8, 11e21.
Belk, K.E., 2012. Quantifying the aging response and nutrient composition for Mittal, G.S., Nadulski, R., Barbut, S., Negi, S.C., 1992. Textural profile analysis test
muscles of the beef round. J. Anim. Sci. 90, 996e1007. conditions for meat products. Food Res. Int. 25, 411e417.
Dominguez-Rodrigo, M., Barba, R., 2006. New estimates of tooth mark and per- Mutungi, G., Purslow, P., Warkup, C., 1995. Structural and mechanical changes in
cussion mark frequencies at the FLK Zinj site: the carnivore-hominid-carnivore raw and cooked single porcine muscle-fibers extended to fracture. Meat Sci. 40,
hypothesis falsified. J. Hum. Evol. 50, 170e194. 217e234.
Dominguez-Rodrigo, M., de la Torre, I., De Luque, L., Alcala, L., Mora, R., Ng, A., Waldron, K.W., 1997. Effect of cooking and pre-cooking on cell-wall chem-
Serrallonga, J., Medina, V., 2002. The ST site complex at Peninj, West Lake istry in relation to firmness of carrot tissues. J. Sci. Food Agr. 73, 503e512.
Natron, Tanzania: implications for early hominid behavioural models. Oyen-Tiesma, M., Cook, R.F., 2001. Technique for estimating fracture resistance of
J. Archaeol. Sci. 29, 639e665. cultured neocartilage. J. Mat. Sci. Med. 12, 327e332.
Dominguez-Rodrigo, M., Pickering, T.R., Semaw, S., Rogers, M.J., 2005. Cutmarked Pietrasik, Z., Shand, P.J., 2005. Effects of mechanical treatments and moisture
bones from Pliocene archaeological sites at Gona, Afar, Ethiopia: implications enhancement on the processing characteristics and tenderness of beef semi-
for the function of the world's oldest stone tools. J. Hum. Evol. 48, 109e121. membranosus roasts. Meat Sci. 71, 498e505.
Dominy, N.J., Vogel, E.R., Yeakel, J.D., Constantino, P., Lucas, P.W., 2008. Mechanical Plummer, T., 2004. Flaked stones and old bones: biological and cultural evolution at
properties of plant underground storage organs and implications for dietary the dawn of technology. Am. J. Phys. Anthropol. S39, 118e164.
models of early hominins. Evol. Biol. 35, 159e175. Pobiner, B.L., Rogers, M.J., Monahan, C.M., Harris, J.W., 2008. New evidence for
Duan, X., Zhang, M., Mujumdar, A.S., Wang, R., 2010. Trends in microwave-assisted hominin carcass processing strategies at 1.5 Ma, Koobi Fora, Kenya. J. Hum. Evol.
freeze drying of foods. Dry. Technol. 28, 444e453. 55, 103e130.
Eng, C.M., Lieberman, D.E., Zink, K.D., Peters, M.A., 2013. Bite force and occlusal Purslow, P.P., 1985. The physical basis of meat texture: observations on the fracture
stress production in hominin evolution. Am. J. Phys. Anthropol. 151, 544e557. behavior of cooked bovine M. semitendinosus. Meat Sci. 12, 39e60.
Ennos, A.R., 2012. Solid Biomechanics. Princeton University Press, Princeton. Purslow, P.P., 1991a. Notch-sensitivity of nonlinear materials. J. Mat. Sci. 26,
Finney, E.E., Hall, C.W., 1967. Elastic properties of potatoes. Trans. Am. Soc. Agr. Eng. 4468e4476.
67, 4e8. Purslow, P.P., 1991b. Measuring meat texture and understanding its structural basis.
Gibson, L.J., Ashby, M.F., 1997. Cellular Solids: Structure and Properties, Second In: Vincent, J.F.V., Lillford, P.J. (Eds.), Feeding and the Texture of Food. Cambridge
edition. Cambridge University Press, New York. University Press, Cambridge, pp. 35e56.
Goh, S.M., Charalambides, M.N., Williams, J.G., 2005. Characterization of the Ragir, S., 2000. Diet and food preparation: rethinking early hominid behavior. Evol.
nonlinear viscoelastic constitutive properties of mild cheddar cheese from Anthropol. 9, 153e155.
indentation tests. J. Texture Stud. 36, 459e477. Ragir, S., Rosenberg, M., Tierno, P., 2000. Gut morphology and the avoidance of
Goodall, J., 1986. The Chimpanzees of Gombe: Patterns of Behaviour. Harvard carrion among chimpanzees, baboons, and early hominids. J. Anthropol. Res. 56,
University Press, Massachusetts. 477e512.
Goren-Inbar, N., Alperson, N., Kislev, M.E., Simchoni, O., Melamed, Y., Ben-Nun, A., Roebroeks, W., Villa, P., 2011. On the earliest evidence for habitual use of fire in
Werker, E., 2004. Evidence of hominin control of fire at Gesher Benot Ya'aqov, Europe. Proc. Natl. Acad. Sci. 108, 5209e5214.
Israel. Science 304, 725e727. Semaw, S., Renne, P., Harris, J.W., Feibel, C.S., Bernor, R.L., Fesseha, N., Mowbray, K.,
Greve, C.L., McArdle, R.N., Gohlke, J.R., Labavitch, J.M., 1994a. Impact of heating on 1997. 2.5-million-year-old stone tools from Gona, Ethiopia. Nature 385,
carrot firmness: changes in cell wall components. J. Agric. Food Chem. 42, 333e336.
2900e2906. Semaw, S., Rogers, M.J., Quade, J., Renne, P.R., Butler, R.F., Dominguez-Rodrigo, M.,
Greve, L.C., Shackel, K.A., Ahmadi, H., McArdle, R.N., Gohlke, J.R., Labavitch, J.M., Stout, D., Hart, W.S., Pickering, T., Simpson, S.W., 2003. 2.6-million-year-old
1994b. Impact of heating on carrot firmness: contribution of cellular turgor. stone tools and associated bones from OGS-6 and OGS-7, Gona, Afar, Ethiopia.
J. Agric. Food Chem. 42, 2896e2899. J. Hum. Evol. 45, 169e177.
Hall, K.R.L., Schaller, G.B., 1964. Tool-using behavior of the California sea otter. Sui, Z.Q., Lucas, P.W., Corke, H., 2006. Optimal cooking time of noodles related to
J. Mammal. 45, 287e298. their notch sensitivity. J. Texture Stud. 37, 428e441.
Hatley, T., Kappelman, J., 1980. Bears, pigs, and Plio-Pleistocene hominids: a case for Teaford, M.F., Ungar, P.S., Grine, F.E., 2002. Paleontological evidence for the diets of
the exploitation of below ground food resources. Hum. Ecol 8, 371e387. African Plio-Pleistocene hominins with special reference to early Homo. In:
de Huidobro, F.R., Miguel, E., Blazquez, B., Onega, E., 2005. A comparison between Ungar, P., Teaford, M. (Eds.), Human Diet: Its Origin and Evolution. Bergin and
two methods (Warner-Bratzler and texture profile analysis) for testing either Garvey, Connecticut, pp. 143e166.
raw meat or cooked meat. Meat Sci. 69, 527e536. Thiel, B.L., Donald, A.M., 2000. Microstructural failure mechanisms in cooked and
Keeley, L.H., Toth, N., 1981. Microwear polishes on early stone tools from Koobi-Fora, aged carrots. J. Texture Stud. 31, 437e455.
Kenya. Nature 293, 464e465. Tornberg, E., 2005. Effects of heat on meat proteins: implications on structure and
King, D.A., Dikeman, M.E., Wheeler, T.L., Kastner, C.L., Koohmaraie, M., 2003. Chilling quality of meat products. Meat Sci. 70, 493e508.
and cooking rate effects on some myofibrillar determinants of tenderness of Ungar, P.S., Grine, F.E., Teaford, M.F., 2006. Diet in early Homo: a review of the
beef. J. Anim. Sci. 81, 1473e1481. evidence and a new model of adaptive versatility. A. Rev. Anthropol. 35,
Klein, R.G., 2009. The Human Career: Human Biological and Cultural Origins, Third 209e228.
edition. The University of Chicago Press, Chicago. Van Loocke, M., Lyons, C.G., Simms, C.K., 2006. A validated model of passive muscle
Laden, G., Wrangham, R., 2005. The rise of the hominids as an adaptive shift in in compression. J. Biomech. 39, 2999e3009.
fallback foods: plant underground storage organs (USOs) and australopith or- Vincent, J.F.V., Jeronimidis, G., Khan, A.A., Luyten, H., 1991. The wedge fracture test: a
igins. J. Hum. Evol. 49, 482e498. new method for measurement of food texture. J. Texture Stud. 22, 45e57.
Lahr, M.M., Wright, R.V.S., 1996. The question of robusticity and the relationship Vogel, E.R., van Woerden, J.T., Lucas, P.W., Atmoko, S.S.U., van Schaik, C.P.,
between cranial size and shape in Homo sapiens. J. Hum. Evol. 31, 157e191. Dominy, N.J., 2008. Functional ecology and evolution of hominoid molar enamel
Lewis, G.J., Purslow, P.P., 1989. The strength and stiffness of perimysial connective- thickness: Pan troglodytes schweinfurthii and Pongo pygmaeus wurmbii. J. Hum.
tissue isolated from cooked beef muscle. Meat Sci. 26, 255e269. Evol. 55, 60e74.
166 K.D. Zink et al. / Journal of Human Evolution 77 (2014) 155e166

Willems, M.E.T., Purslow, P.P., 1997. Mechanical and structural characteristics of Wrangham, R., 2007. The cooking enigma. In: Ungar, P. (Ed.), Evolution of the Hu-
single muscle fibres and fibre groups from raw and cooked pork longissimus man Diet: the Known, the Unknown, and the Unknowable. Oxford University
muscle. Meat Sci. 46, 285e301. Press, New York, pp. 308e323.
Wolpoff, M.H., 1971. Metric Trends in Hominid Dental Evolution. Press of Case Wrangham, R., Conklin-Brittain, N., 2003. 'Cooking as a biological trait'. Comp.
Western Reserve University, Cleveland. Biochem. Physiol. A 136, 35e46.
Wolpoff, M.H., 1973. Posterior tooth size, body size, and diet in South African gracile Wrangham, R.W., Jones, J.H., Laden, G., Pilbeam, D., Conklin-Brittain, N., 1999. The
australopithecines. Am. J. Phys. Anthropol. 39, 375e393. raw and the stolen: cooking and the ecology of human origins. Curr. Anthropol.
Wood, B., Collard, M., 1999. The changing face of genus Homo. Evol. Anthropol. 8, 40, 567e594.
195e207.

You might also like