You are on page 1of 20

Downloaded from http://rsta.royalsocietypublishing.

org/ on May 27, 2015

Influence of structural
hierarchy on the fracture
rsta.royalsocietypublishing.org
behaviour of tooth enamel
Ezgi D. Yilmaz1 , Gerold A. Schneider1
Review and Michael V. Swain2,3
1 Institute of Advanced Ceramics, Hamburg University of Technology,
Cite this article: Yilmaz ED, Schneider GA,
Swain MV. 2015 Influence of structural Hamburg, Germany
hierarchy on the fracture behaviour of tooth 2 Prosthetic Department, Freiburg University, 29106 Freiburg,

enamel. Phil. Trans. R. Soc. A 373: 20140130. Germany


http://dx.doi.org/10.1098/rsta.2014.0130 3 Biomaterials, Faculty of Dentistry, University of Sydney, Sydney,

New South Wales 2006, Australia


One contribution of 19 to a theme issue
‘Fracturing across the multi-scales Tooth enamel has the critical role of enabling the
of diverse materials’. mastication of food and also of protecting the
underlying vital dentin and pulp structure. Unlike
most vital tissue, enamel has no ability to repair
or remodel and as such has had to develop robust
Subject Areas: damage tolerance to withstand contact fatigue events
materials science throughout the lifetime of a species. To achieve such
behaviour, enamel has evolved a complex hierarchical
Keywords: structure that varies slightly between different species.
enamel, hydroxyapatite, hierarchy, fracture The major component of enamel is apatite in
the form of crystallite fibres with a nanometre-
toughness, R-curve, bending strength
sized diameter that extend from the dentin–enamel
junction to the oral surface. These crystallites are
Author for correspondence: bound together by proteins and peptides into a
Michael V. Swain range of hierarchical structures from micrometre
e-mail: michael.swain@sydney.edu.au diameter prisms to 50–100 µm diameter bundles
of prisms known as Hunter–Schreger bands. As a
consequence of such complex structural organization,
the damage tolerance of enamel increases through
various toughening mechanisms in the hierarchy
but at the expense of fracture strength. This review
critically evaluates the role of hierarchy on the
development of the R-curve and the stress–strain
behaviour. It attempts to identify and quantify the
multiple mechanisms responsible for this behaviour
as well as their impact on damage tolerance.

Electronic supplementary material is available


at http://dx.doi.org/10.1098/rsta.2014.0130 or
1. Introduction
via http://rsta.royalsocietypublishing.org. Enamel is the most highly mineralized tissue (up to 96%
by volume [1]), and thus the hardest material in the

2015 The Author(s) Published by the Royal Society. All rights reserved.
Downloaded from http://rsta.royalsocietypublishing.org/ on May 27, 2015

hierarchical order of enamel according to Koenigswald and Clemen


2
level 4 level 3 level 2 level 1
schmelzmuster enamel types multiple prisms multiple crystallites

rsta.royalsocietypublishing.org Phil. Trans. R. Soc. A 373: 20140130


.........................................................
(enamel pattern)

200 mm
sin
ba gle
nd 20 mm level 0
decussating enamel single crystallite

2 mm 200 mm
outer inner
enamel enamel dentin prism

dentin–enamel junction 20 mm interprismatic matrix


radial enamel

Figure 1. Microstructure of bovine enamel and its hierarchical levels defined according to Koenigswald & Clemen [14,15]:
a single nanometre-sized hydroxyapatite (HAP) crystallite (level 0) is the elementary building block of enamel. Multiple
crystallites are characterized as level 1. Prisms (p) and ipm formed by multiple HAP crystallites are the micrometre-scaled
structural elements at the next level of hierarchy (level 2). The arrangement of prisms defines the next level of hierarchy, enamel
types (level 3). Prisms with a parallel arrangement form ‘radial enamel’. Prism bundles crossing each other build decussating
enamel. Prism bundles forming well-defined bands, and decussating in a periodic fashion, are a subtype of decussating enamel
specified as Hunter–Schreger bands (HSBs). Within each band, prisms have the same orientation, which differs from the
neighbouring band. Spatial distribution of enamel types in a single tooth determines the schmelzmuster (enamel pattern;
level 4). In most mammals, the inner part of the enamel is occupied by decussating enamel types, and the outer region is
made of radial enamel. (The first image on the left is reproduced with permission from [16].) (Online version in colour.)

human body, and is located in the outermost layer of the crown of a tooth. Its biological function
is to protect the underlying soft dentin and pulp from mechanical forces and chemical attacks
during tooth function. Enamel is not a living tissue and thus its loss owing to abrasion and
fracture is irrecoverable. This tissue has frequently been described as a brittle material in earlier
studies [2–4] and thus relatively little attention has been given to its structure–fracture behaviour
relations in comparison with those of bone and dentin. On the other hand, it was revealed that,
in spite of pre-existence of cracks in the outer surface [5], enamel can survive millions of chewing
cycles with mastication forces ranging between 28 and 1200 N [6]. This raised the question of
how an inherently brittle tissue sustains cracks without catastrophic fracture [5]. Identifying
the features that enable this performance in biological materials with the aim of implementing
them to improve engineering materials has become an increasingly important area over many
years. This has resulted in renewed interest with further enamel studies employing advanced
fracture mechanics being published [7–13]. These authors reported that bulk enamel exhibits
increasing fracture toughness with crack extension (R-curve behaviour) compared with brittle
materials that generally have a single value of fracture toughness. This behaviour of enamel is
attributed to the interaction of cracks with its hierarchical microstructure, which will be classified
in this review according to Koenigswald and Clemen’s structural description [14,15]. The basic
structural element in enamel is a single hydroxyapatite (HAP) crystallite (level 0). Bundles of
these crystallites (level 1) form the micro-scale building blocks known as prisms (p) and the
interprismatic matrix (ipm; level 2; the term ‘prism’ is used synonymously with ‘rod’). The
arrangement and orientation of prisms determines the next level of hierarchy termed as enamel
types (level 3), which can be broadly classified as radial and decussating enamel. The next level
of hierarchy (level 4) is called the ‘schmelzmuster’ (enamel pattern), which defines the three-
dimensional distribution of the different enamel types in a tooth. As a typical representation, an
overview of the hierarchical organization in bovine enamel is illustrated in figure 1. In addition,
enamel is known to contain a limited amount of soft organic matter (protein/peptides, firmly
Downloaded from http://rsta.royalsocietypublishing.org/ on May 27, 2015

(a) (b) hierarchical strength for enamel 3


10 000
quasi-self-similar hierarchical HAP strength

rsta.royalsocietypublishing.org Phil. Trans. R. Soc. A 373: 20140130


.........................................................
strength (MPa)
structure loaded in
consistency
tension of model and
1000 calculation
level n

level 2 100
level 1 0 1 2 3
no. hierarchy

(c) hierarchical elastic modulus for enamel


90

elastic modulus (GPa)


80 EHAP =
70
60 80 GPa
50
40
bottom-up design 30
20 Gprotein =
10 300 MPa
0
0 1 2 3
no. hierarchy

Figure 2. Hierarchical mechanics model (image from Bechtle et al. [16]) and supporting experimental results: quasi-self-similar
brick and mortar hierarchical model loaded in tension according to Gao and co-workers [28] (a). The brick represents hard
particles (black), which are enveloped by mortar, the soft phase (grey), at each level. The composite structure at level n forms
the hard particles at level (n + 1). One basic assumption of this model is that hard particles carry the entire tensile load. The
soft phase transfers the tensile load via shear to the particles. The reduction in strength (b) and elastic modulus (c) in bovine
enamel with increasing hierarchy was experimentally assessed by Bechtle et al. [16] and applied to verify the model proposed
by Gao and co-workers [28]. The orange, green and blue regions represent, respectively, the experimental data, properties of
basic constituents and calculations using Gao’s model.

and loosely bound water) [17–24], which accumulates at sites where mineral density is low such
as intercrystallite and interprismatic spaces [16]. Details regarding the microstructural features
are described later in this review. It should be noted here that there exists no strict definition
regarding the counting of hierarchical levels. We adopted Koenigswald and Clemen’s hierarchical
terminology here as it is well suited to the diversity in enamel microstructure that has to be
taken into account among different species. However, according to the hierarchical classification
approach based on size, there is no difference between ‘multiple prisms’ and ‘radial enamel’, both
are considered as level 2 [9,11,16,25,26].
The combination of being hard, stiff and tough—as observed generally in mineralized
biological tissues—is a desired mechanical property also for engineering materials and hence
much recent effort has been directed at exploring the basic mechanical principles associated
with these hierarchical microstructures both experimentally and theoretically. Novel experimental
techniques have enabled researchers to assess the individual response of microstructures over
different length scales and constituent materials by combining nano- to macro-scale mechanical
tests together with high-resolution microscopes and synchrotron X-ray diffraction. A considerable
amount of literature has been published on bone identifying its hierarchical deformation
mechanisms, as summarized by Gupta & Zioupos [27]. This helped form the basis for the
hierarchical mechanics model developed by Gao and co-workers [28] based on suggestions from
Downloaded from http://rsta.royalsocietypublishing.org/ on May 27, 2015

Jäger & Fratzl [29]. Furthermore, by applying fracture mechanics concepts to a two-dimensional,
4
quasi-self-similar structure resembling bone, they derived formulae for mechanical properties,
including strength and elastic modulus for higher structural levels as functions of the material

rsta.royalsocietypublishing.org Phil. Trans. R. Soc. A 373: 20140130


.........................................................
properties of subordinate hierarchical levels (e.g. protein content, mineral particle size or particle
strength). The reader is referred to recent publications [28–33] for more details.
The basic concept of the model is visualized in figure 2a. The quasi-self-similar structure is
considered to be built of hard mineral particles enveloped by a soft protein matrix at the lowest
hierarchical level. This composite structure at level n further forms hard particles at level (n + 1)
which again are enveloped by soft matrix. This means that each additional level of hierarchy
adds more protein to the composite structure (assuming that the mineral volume fraction of each
hierarchy is constant), which improves damage tolerance at the expense of the elastic modulus
and strength under tension as experimentally observed in enamel from level 1 to level 2 by Bechtle
et al. [16] (figure 2b,c). Bechtle et al. analysed these experimental results with Gao’s model and they
were found to be in good agreement (figure 2b,c).
Besides measurements of stress–strain curves at different length scales, there have been
contributions in enamel research investigating the submicrometre fracture toughness and
toughening mechanisms with small-scale testing methods [25,26]. Apart from these, additional
R-curve measurements provided greater insight into the orientation-related factors associated
with crack growth resistance of enamel at larger scales. The objective of this review is to bring
together the findings reported in different studies associated with different hierarchical levels
so as to provide a deeper understanding of the role of hierarchy in the fracture behaviour of
enamel. This requires a detailed description of its microstructural features. Thus, we initially
summarize the structural features at individual hierarchical levels, and then analyse the role
and contribution of the corresponding features to the fracture behaviour (focusing on fracture
toughness, fracture strength, deformation curves, toughening mechanisms). An overview of the
characterization methods used in the literature [2–13,16,25,26,34–36] that are reviewed here for
assessing the fracture behaviour of enamel at different length scales can be found in the electronic
supplementary material. Finally, we briefly summarize the findings.

2. Structural characteristics of enamel at different hierarchical levels


and their role on fracture behaviour
(a) Hierarchical level 0: single crystallite
(i) Structural features
The hierarchical level 0 of enamel is represented by a single crystallite (figure 1). The reported
values of the cross-sectional dimensions of the mature crystallites from various sources were listed
by Daculsi & Kerebel [37], and they ranged between 15 and 50 nm in thickness, and between 40
and 150 nm in width. The length of the crystallites was shown to be at least 100 µm [38]; however,
they are believed to extend without interruption from the dentin–enamel junction (DEJ) to the
outer enamel surface (OES) [39]. It should also be mentioned here that the crystallites of enamel
are a carbonated form of HAP that differs slightly from synthetic HAP [39]. Additionally, HAP
crystallites in enamel are derived from a biomineralization process which results in imperfections
as evidenced by high-resolution electron microscopy [37,40]. Enamel is formed by ameloblast cells
which secrete the matrix proteins that are responsible for deposition of the mineral crystallites
[39]. Matrix proteins self-assemble to form nanospheres surrounding the crystallites and force
them to grow along their long axis and inhibit crystal–crystal fusion. During maturation, proteins
break down and are removed from the matrix to create space, so that the pre-existing crystallites
can grow in width and thickness. This allows the crystallites to come into contact and may induce
mechanical stresses distorting the atomic arrangement and forming dislocations. Moreover, the
external surfaces of the crystallites are not smooth but display frequent step-like irregularities
Downloaded from http://rsta.royalsocietypublishing.org/ on May 27, 2015

[41]. The influence of the dimensions and imperfections of the enamel crystallites is analysed in
5
the following.

rsta.royalsocietypublishing.org Phil. Trans. R. Soc. A 373: 20140130


.........................................................
(ii) Implications for fracture behaviour
The specific nanometre-sized scale of the enamel HAP crystallites has evolved by nature as
postulated by Gao et al. [30–33] in order to ensure optimum fracture strength of the mineral
particles. Brittle materials fail by stress concentrations that arise owing to the presence of crack-
like flaws before reaching their theoretical strength σth , which is the tensile stress necessary to
break primary chemical bonds in a defect-free material. According to Griffith’s theory, fracture
stress depends on the material parameters elastic modulus E and surface energy γ and on the
crack size a. If the dimension of the biomineral crystallites drops below a critical size, they become
defect insensitive and the atomic bonds control the strength [30]. The critical length scale for defect
tolerance, a∗ , is estimated as
γE
a∗ ≈ α 2 , (2.1)
σth

where α is a parameter dependent upon crack geometry and is approximately equal to π for
a half-cracked platelet [30]. By taking the parameters γHAP = 1 J m−2 [42], EHAP = 125 GPa [43]
and σth = E/30 [30], a∗ is estimated to be approximately 36 nm. This is in agreement with the
dimensions of the HAP crystallites in enamel and favours the concept of the optimization of
the fracture strength of the mineral crystallites and their insensitivity to minor defects. This
has not been validated owing to the lack of experimental techniques enabling the assessment
of the fracture strength of a single HAP crystallite isolated from a biological tissue. However,
Bechtle et al. [16] calculated the fracture strength of a single enamel crystallite to be approximately
2 GPa indirectly, by combining the experimental results and analytical solutions for hierarchical
materials. This value is approximately one order of magnitude higher than the strength obtained
from bulk HAP (approx. 100 MPa [44]; see the electronic supplementary material, table S1) and
is of the order of theoretical strength estimation of approximately 4 GPa (σth = E/30). The slightly
lower experimental value compared with the theoretical value may be explained by the fact that
the elastic modulus of a single enamel crystallite is not reported in the literature owing to the lack
of sample preparation and testing technique. Thus, the elastic modulus of an artificial micro-sized
HAP single crystal (125 GPa) assessed by nanoindentation [43] is used in the calculation of the
theoretical strength. Note that indentation approaches quantify material properties under multi-
axial stress state (largely compressive stresses) that can differ from the properties assessed under
uniaxial tension. Considering the existence of the aforementioned imperfections in the enamel
crystallites, the tensile elastic modulus is expected to be lower in reality than the elastic modulus
value given here. This could be the reason for the discrepancy between the estimated theoretical
strength here and the reported fracture strength by Bechtle et al. [16].

(b) Hierarchical level 1: multiple crystallites


(i) Structural features
The hierarchical level 1 of enamel is associated with bundles of crystallites held together through
the composite nature of enamel and imperfections in the crystallite shapes and arrangements,
which will be elucidated in this section. Enamel is formed in two stages as explained above.
During the initial stage, the crystallites are thin and ribbon-like structures whose unit cell is
hexagonal (6/m). However, the hexagonal outline of the crystallites is lost as they press against
each other during the final phase of their growth and become irregular in shape [37,39,45]. These
interlocking shapes grow to fill all the available space between the crystallites [46] and hold them
together (figure 3a). In addition, some other structural abnormalities such as crystallite fusion
and lateral branches (bifurcations) are also evident in enamel crystallites [37,45]. Examining
longitudinal sections of mature enamel reveals that HAP crystallites do not align perfectly parallel
Downloaded from http://rsta.royalsocietypublishing.org/ on May 27, 2015

(a) 6

rsta.royalsocietypublishing.org Phil. Trans. R. Soc. A 373: 20140130


.........................................................
200 nm

(b) (c)

1 mm 500 nm

Figure 3. Cross sections of HAP crystallites of mature rat incisor enamel obtained by transmission electron microscopy (a)
(reproduced with permission from [45]), and a longitudinal view of HAP crystallites in bovine enamel revealing their entangled
arrangement obtained by scanning electron microscope (b,c). (Online version in colour.)

to each other; they seem rather entangled (figure 3b,c). This imperfect arrangement may interlock
the crystallites, enabling them to remain attached structurally. However, a more common view
is that the crystallites are bound together by intercrystallite proteins, which are not incorporated
within the crystalline HAP. It is observed that there exist remnant proteins and water in mature
enamel as assessed by vibrational spectroscopy (infrared and Raman) and thermogravimetric
analysis. These soft substances are postulated to fill the free space where the mineral density is
low owing to imperfect crystal packing. These regions, which are 1–2.5 nm electron-lucent spaces
between the crystallites, are detected frequently in transmission electron microscopic (TEM)
studies [41,45]. However, it is also observed that the crystallites appear to be in direct contact with
each other. This indicates that there may be protein layers between crystallites but they are not
homogeneously distributed. The type of intercrystallite proteins in mature enamel was reported
as amelogenin, which has a hydrophobic character, the significance of which will be described
below [47,48].

(ii) Implications for fracture behaviour


Characterization of the mechanical properties of the constituents at level 1 solely, without any
contribution from higher-order structural motifs and materials, has only a recent history. The
first study on the small-scale mechanical characterization of enamel was published by Chan et al.
[49] using a micro-cantilever bending technique. Following this study, several others used the
same methodology to investigate the properties of isolated crystallites from different tooth types
[16,50]. The resultant mechanical properties from the bundles of crystallites l (loaded parallel to
the crystallite long axis) from various studies are listed in the electronic supplementary material,
table S1. The micrographs of the samples before and after loading together with the resultant
stress–strain curves are illustrated in figure 4a,b. Chan et al. [50] investigated the influence of
dehydration owing to focused ion beam (FIB) treatment and could not find a significant effect.
Although there is substantial scatter in the failure strengths obtained from different studies (that
can be attributed to different mineral content of the tested samples), the fracture strength of this
Downloaded from http://rsta.royalsocietypublishing.org/ on May 27, 2015

level is approximately 1 GPa. This is still one order of magnitude higher than bulk HAP strength
7
(approx. 100 MPa [44]) but two times lower than the strength of a single HAP crystallite (approx.
2 GPa [16]). This difference can be explained by the composite nature of the first hierarchical

rsta.royalsocietypublishing.org Phil. Trans. R. Soc. A 373: 20140130


.........................................................
level, as proteins are also incorporated. Moreover, there is a higher probability of the presence
of defects and imperfections in a bundle of crystallites. The elastic modulus values measured via
micro-cantilever bending experiments are significantly lower than the elastic modulus of bulk
HAP (approx. 80 GPa [44]) assessed under bending. The possibility of degradation in the HAP
crystallites owing to FIB treatment was elaborated using TEM [51,52], where HAP is reported to
be structurally resistant to FIB-induced damage. Moreover, Chan et al. [49] revealed no significant
difference in the elastic modulus and hardness of enamel before and after FIB preparation,
which would not be expected in the presence of protein degradation [53]. Consequently, it can
be assumed that FIB treatment does not have a significant effect on the measured mechanical
properties of enamel. The resulting low elastic modulus values were attributed to the dissolution
of minerals in the micro-cantilevers by the acid treatment (this etching procedure is necessary
to make the prism boundaries visible and locate the milling site correctly) [49]. Moreover, as
mentioned earlier, the HAP crystallites differ slightly from synthetic HAP in chemistry and
structure, which may also alter their mechanical properties [35].
In soft-matrix-based structural composites, progressive failure behaviour (saw-tooth-shaped
stress–strain curve) is expected to occur, where the reinforcing fibres fail in sequence beginning
with the weakest [54,55]. As each succeeding fibre fails, the remaining intact fibres share the
increased load released by the broken fibres that is distributed by the matrix [55]. However, the
commonly observed fracture behaviour at this hierarchical level was catastrophic failure after a
near entirely linear–elastic deformation as shown in figure 4b. There is a hint of slight nonlinear
behaviour from beyond two-thirds of the maximum stress, but whether this is associated with
progressive crystallite failure or shear between the crystallites is unclear. Alternatively, the
catastrophic failure indicates that all the HAP crystallites in the micro-cantilever bending test
failed simultaneously. This process may be attributed to the lack of load transfer ability of
the intercrystallite proteins owing to their limited content (less than 2 vol.%) compared with
the matrix content in the above-mentioned artificial composites (greater than 30 vol.% [55]). The
fracture toughness in this orientation, where the crystallites are parallel to the tensile stress
direction (crystallite fracture mode), was assessed by Bechtle et al. [26] by using notched micro-
cantilevers as schematically illustrated in the electronic supplementary material, figure S1. The
notch placed with an off-set from the cantilever’s fixed ends introduced a shear component KII
at the crack tip in addition to the tensile component KI [26]. KI and KII acted simultaneously at
the crack tip when fracture occurred but the latter was found to be very small (see the electronic
supplementary material, table S2). Upon computing the corresponding geometry function for
the triangular notched sample geometry, Bechtle et al. [26] assessed the KIC values as 0.44–
0.91 MPa m0.5 , which lies in the range of synthetic HAP values, indicating the minor role of
intercrystallite proteins in fracture toughness in this orientation. However, when the crack
propagation is induced between the crystallites, the fracture toughness values of bovine enamel
showed a much larger scatter [25]. Ang et al. [25] measured the crack opening displacement
at the crack tip region induced by Vickers indentation and evaluated both mode I KI0 and
mode III KIII0 crack tip toughness (which is the toughness at crack initiation, where no extrinsic
toughening mechanisms are yet activated) by applying the cohesive zone solution of the
Dugdale–Muskhelishvili crack model and Irwin’s ‘near-field’ solution. The cracks propagated
mainly along the prism boundaries (considered as hierarchical level 2 in this review) and within
the prisms, the latter corresponding to hierarchical level 1. The values of crack tip toughness
KI0 and KIII0 ranged between 0.52 and 1.62 MPa m0.5 and between 0.05 and 0.15 MPa m0.5 ,
respectively [25]. The authors also computed the crack tip closure stresses as 434–779 MPa for
a cohesive zone width of 24–36 nm and a length of 1.6–3.2 µm, which may be attributed to
microcracking, bridging by protein filaments and crystallite ligaments [25]. The last are the
toughening mechanisms prevailing at the first hierarchical level of enamel as illustrated in
figure 5. Microcracks of 100–500 nm in length were observed around the main crack within an
Downloaded from http://rsta.royalsocietypublishing.org/ on May 27, 2015

hierarchical level 1—multiple crystallites


8
(a) (b)
1000

rsta.royalsocietypublishing.org Phil. Trans. R. Soc. A 373: 20140130


.........................................................
fracture
750

stress (MPa)
500

250

0
0 0.5 1.0 1.5 2.0
3 mm
strain (%)
hierarchical level 2—multiple prisms
(c) (d)
400

fracture
300

stress (MPa) 200

100

0
0 0.4 0.8 1.2 1.6
10 mm
strain (%)

Figure 4. Micro-cantilever bending tests. (Images from Bechtle et al. [16].) (a) Triangular cross-section micrometre-sized
(<5 µm in width and height, approx. 15 µm in length) FIB-milled cantilevers within the prisms avoiding the prism sheaths.
These samples consisted of HAP crystallites only, representing the hierarchical level 1. Loading was achieved by using the optical
microscope of the nanoindentation system, and the Berkovich tip was positioned at the cantilever’s free end, as marked by
the arrow. A typical stress–strain curve of HAP crystallites loaded parallel to their long axis is displayed in (b). The inset in (b)
displays the post-failure morphology, where the broken crystallites protrude from the fracture surface. (c) For characterization
of the second hierarchical level in enamel, larger cantilevers (10–15 µm in width and height, 40–50 µm in length) covering
several prisms were prepared and tested with the same method. A representative stress–strain curve for the multiple prisms is
demonstrated in (d). Note the increased nonlinearity in the stress–strain curve compared with level 1 that is attributed to the
increased protein content with increasing hierarchical levels [16]. The fracture mode of the multiple prisms is shown in the inset
(d), revealing the crystallite fracture within the prisms together with the crack deflection at prism boundaries that run parallel
to the length of the beam. (Online version in colour.)

enamel prism and followed the contours of the crystallites (figure 5a). These microcracks arise
as the main crack is obstructed by the frictional mineral–mineral interactions associated with the
irregular cross-sectional shapes of the crystallites. The latter behaviour was observed by Ang et al.
[25] when they examined the crack profiles using atomic force microscopy (AFM) and identified
nano-sized approximately 85 nm frictional bridges upon scanning close to the crack tip as shown
in figure 5c. In addition, unbroken crystallite ligaments bridging the main crack propagating
across the long axis of the crystallites are demonstrated in figure 5a. The entangling arrangement
of the crystallites together with the bifurcations and fusions within them complicates the crack
propagation along the crystallite axis by the formation of crystallite bridges as can be seen in
a TEM cross section (figure 5b) showing a crack induced by nanoindentation forming around
the long axis of the crystallites [56]. Aside from the observations of microcracks and crystallite
bridges, both the studies identified the presence of protein ligament bridges when the crack
propagates between crystallites.
Downloaded from http://rsta.royalsocietypublishing.org/ on May 27, 2015

(a) an across (^) crack (b) protein ligament 9


seemingly
protein

rsta.royalsocietypublishing.org Phil. Trans. R. Soc. A 373: 20140130


.........................................................
ligaments crack

ck on
cra gati
a ~500 nm
op
pr microcrack crystallite
ligament
~100 nm 200 nm
crystallite
ligament microcracks
50 nm

(c) (an across (^) crack)

crack
~85 nm
propagation
frictional
bridge
crack 200 nm

Figure 5. Toughening mechanisms at hierarchical level 1: all images show the crystallite separation mode that is captured
within a prism, in which the transversal (a,c) and longitudinal cross sections (b) of crystallites are observed. (a) Microcracks
around the main crack induced by Vickers indentation in bovine enamel and seemingly protein and crystallite ligaments bridging
the main crack are detected by SEM. (Reproduced with permission from [25].) (b) TEM investigation following a 500 mN
Berkovich indent on human enamel also revealed protein and crystallite crack bridges directly beneath the indenter. (Image
from Xie et al. [56].) (c) An AFM topography image showing a frictional bridge corresponding to the length scale of one HAP
crystallite. (Reproduced with permission from [25].) (Online version in colour.)

(c) Hierarchical level 2: multiple prisms


(i) Structural features
Enamel microstructure is a result of millions of years of evolution and it has diversified
among different species and to a lesser extent between the teeth of an individual. Evolutionary
development of enamel has an adaptive relationship to the stresses generated during the
functioning of the tooth and is closely related to the physiological task of the dentition and
the properties of the food items [57]. The enamel of lower vertebrates has the simplest structure
known as ‘prismless enamel (parallel crystallite enamel)’, representing the most primitive enamel
type with only one level of hierarchy (multiple crystallites). Prismless enamel is observed in the
teeth of species (mainly reptiles) whose physiological task is usually limited to the capturing,
piercing and holding of the prey with no contact between the opposing jaws [58,59]. The change
in feeding habits to mastication and chewing resulting in higher stresses generated in teeth was
accompanied by a modification of enamel microstructure from prismless to prismatic enamel
in mammals [15,58]. Prismatic enamel evolved from prismless enamel as a consequence of the
differentiation in the enamel formation (ameloblast) cells [58], which promoted the creation of gap
regions between the bundles of crystallites forming enamel prism (rods; figure 6). Each ameloblast
cell is responsible for the formation of a single prism (p) and also some ipm component, which is
also associated with the hierarchical level 2. Enamel prisms are continuous from the DEJ close
to the OES and have roughly cylindrical geometry but their exact shape (round, angled, arc-
shaped, keyhole-shaped in cross section) may vary taxonomically and/or locally in a single tooth.
Roughly, the thickness/width/diameter (depending on the shape) of prisms ranges from 1 to
Downloaded from http://rsta.royalsocietypublishing.org/ on May 27, 2015

(a) (b) (c) 10

rsta.royalsocietypublishing.org Phil. Trans. R. Soc. A 373: 20140130


.........................................................
1 mm 1 mm 1 mm

Figure 6. Closed prism boundaries (a) and arc-shaped open prism boundaries in bovine enamel (b). Keyhole-shaped open
prism boundaries in human enamel (c) that are surrounded and laminated by interprismatic matrix. The shape of the prism
boundaries is marked with blue lines. (Online version in colour.)

7 µm, and the thickness of the ipm can vary between 0.5 and 3 µm depending on the enamel type.
Crystallites within a prism are oriented roughly parallel to the prism long axis; however, in some
species (such as the keyhole-shaped prisms in humans, as seen in figure 6c), they fan out from
the centre towards the edges [15]. Crystallites within the ipm are parallel to one another but at an
angle to the prism direction in general, whose variations will be addressed in the explanation of
prism patterns.
Prisms have a borderline called a ‘prism boundary’ that is a gap zone, whose thickness appears
to be 0.1–1 µm in the micrographs shown in figure 6. However, this can be misleading, because
those samples were highly acid etched prior to microscopic inspection to make the boundaries
visible. In the TEM micrographs of unetched samples, the prism boundaries are less prominent
and are revealed as discontinuous regions with small pores (less than 50 nm) partially bridged
by crystallites [56,60]. Prism boundaries have been proposed to be made up of a protein sheath
[39], because these are the gap regions where the majority of the remnant proteins should be
constricted during maturation. However, it has been almost impossible to observe the protein
sheaths directly by structural analysis owing to the degradation of proteins by harsh sample
preparation procedures. To date, there is only one report [61] demonstrating the protein sheaths
(see the electronic supplementary material, figure S2). This study enabled researchers for the
first time to identify the nature of enamel proteins as a three-dimensional fibrous mesh-like
structure. The interprismatic proteins in mature enamel were reported to be ameloblastin and
are hydrophilic [62]. The hydrophobic nature of amelogenin versus the hydrophilic nature of
ameloblastin enables aqueous solutions to penetrate between the prisms but less so between the
crystallites. For human enamel, this has a significant influence on the effects of bleaching agents
penetrating teeth [63].
The appearance of prism boundaries and differences in prism–ipm packing (or the way the
prisms and ipm are arranged relative to another) designated as prism patterns are the most
striking characteristics at this level that vary among mammalian taxa [64]. The prisms may have
‘closed’ boundaries, surrounding the prisms completely or ‘open’ (incomplete) boundaries, where
the ipm anastomoses to prisms and forms a single continuum as shown in figure 6. Many species
show a combination of both [15]. The arrangement of prisms and ipm can be simply distinguished
between honeycomb and sandwich patterns (figure 7). In the honeycomb pattern, the ipm
surrounds the prisms that are organized roughly in a disordered way. In the primitive condition,
the orientation of the crystallites in the ipm remains more or less parallel with that of the
crystallites of prisms, which has however diverged with time and the intersecting angle between
the crystallites of the ipm and prisms has increased [65]. In the sandwich pattern, prisms are
arranged in approximately vertical rows separated by the ipm layers (in these arrangements the
ipm is also called ‘inter-row sheets’). The ipm layers have bifurcations that create imperfections
in the arrays as shown in figure 7a. The crystallites of the prisms and ipm intersect at almost right
angles in the sandwich pattern. This is also known as ‘crystallite decussation’ and is believed to
emerge at a more derivative stage of enamel evolution as a fracture-resisting mechanism at this
hierarchical level [65].
Downloaded from http://rsta.royalsocietypublishing.org/ on May 27, 2015

(a) (b) 11

rsta.royalsocietypublishing.org Phil. Trans. R. Soc. A 373: 20140130


.........................................................
5 mm 5 mm

Figure 7. Sandwich pattern of prism rows and ipm layers (inter-row sheets; a); honeycomb pattern of prisms surrounded by
ipm (b) in bovine enamel. Bifurcations in the ipm layers are marked with the arrows. (Online version in colour.)

(ii) Implications for fracture behaviour


The fracture strength and elastic modulus of several prisms isolated from the bulk enamel
(figure 4c,d) were reported to be 478 ± 93 MPa and 36 ± 8 GPa, respectively [16]. There was almost
a 50% decrease in the fracture strength and elastic modulus values compared with level 1, which
was attributed to the possibly higher protein content of level 2 by Bechtle et al. according to Gao’s
hierarchical model, as introduced previously [16]. Gao calculated the overall mineral content,
ϕtotal , of hierarchical materials with the following relationship:

ϕtotal = ϕ1 ϕ2 . . . ϕn , (2.2)

where ϕn is the mineral volume fraction of the composite structure at level n [31]. Assuming the
mineral volume fraction of each hierarchical level is constant (ϕ1 = ϕ2 = ϕn ), the mineral volume
fraction decreases with increasing hierarchical structuring. In addition, the nonlinear character
of the stress–strain curves appears to have increased from level 1 to level 2 (figure 4d). The
subtle chemical differences between the interprismatic proteins and the intercrystallite proteins
were previously mentioned. The former are hydrophilic, whereas the latter are hydrophobic.
Consequently, the increased nonlinearity at level 2 could be related to the higher deformability
of the more porous water-filled protein structure at the prism boundaries compared with the
limited water present within the prisms. Furthermore, the crystallite fracture within the prisms is
accompanied by crack deflection at the prism boundaries, indicating also the change in fracture
mode at this level. The KIC results assessed by FIB-notched micro-cantilever bending were 0.83–
0.89 MPa m0.5 [26]. The mode I crack tip fracture toughness values estimated by Irwin’s crack
tip solution for the second hierarchical level of enamel are 0.64–0.94 MPa m0.5 for the cracks
propagating along the prism boundaries [25]. Results from the earlier Vickers indentation studies
on human enamel reporting the fracture toughness are also associated here with the hierarchical
level 2 owing to the corresponding crack lengths (approx. 50 µm). These values ranged from
0.68 to 1.3 MPa m0.5 . Correspondingly, crack closure stresses near the crack tip along the prism
boundaries were estimated as 168–754 MPa for a cohesive zone length of 10–1.6 µm [25].
Evolution of prismatic enamel from prismless enamel has resulted in the formation of prism
boundaries that introduced a weak path for crack propagation. This serves in the first place to
localize the cracks into confined zones, which control the crack path direction. This guided crack
growth, resulting from various crack-arresting mechanisms, contains the crack within the tooth
and so prevents chipping of enamel [8]. These mechanisms associated with level 2 are represented
in figure 8. Note that some of these mechanisms are detected only in specific parts of the tissue,
which will be addressed at level 4 of the schmelzmuster. It was previously mentioned that the
high angle (approx. 90◦ ) between the prisms and ipm sheets (sandwich pattern) creates a prism–
ipm (crystallite) decussation, which differs from the decussation of prism bundles (HSBs). This
Downloaded from http://rsta.royalsocietypublishing.org/ on May 27, 2015

crack deflection crack branching microcracking


12
meandering k n
(a) (b) (c) ac io
crack cr gat
a
op

rsta.royalsocietypublishing.org Phil. Trans. R. Soc. A 373: 20140130


.........................................................
pr

crack
branching

4 mm 5 mm 5 mm

bridging by protein sheaths bridging by ipm bridging by multiple prisms


(d) (e) (f)

ipm

5 mm 2 mm 30 mm

Figure 8. Toughening mechanisms at the hierarchical level 2: crack deflection owing to prism–ipm decussation in bovine
enamel (a), crack branching and meandering in bovine enamel (reproduced with permission from [11]) (b), microcracks about
the main crack identified at prism boundaries and ipm regions in bovine enamel (c), protein ligaments of a few micrometres
thickness bridging a crack propagating through prism boundaries evident on the crack face of human enamel (reproduced with
permission from [7]) (d), unbroken mineral ligaments consisting of ipm (e) and prism bundles (f ) identified in bovine enamel.
(Online version in colour.)

is commonly observed in ungulates and here the crack deflection occurs at right angles owing
to prism–ipm decussation evident in bovine enamel as presented in figure 8a. Such deflections
result in a decrease in the local stress intensity at the crack tip [7]. In addition, some of the ipm
sheets and the prisms have to be broken for the crack to advance, manifesting the reinforcing
strategy of enamel at this level of hierarchy. Uncracked ligament bridging of proteins residing
between the prisms is evident from observations of the crack path in both human and bovine
enamel as an energy-dissipating mechanism through extension/unfolding of the peptide chains
[7–9] and is illustrated in figure 8d. Open prism boundaries where the ipm anastomoses to prisms
at the tail region forming a single continuum were also introduced previously in the structural
features section. These crystallite interconnections between the prisms and ipm are responsible for
crack bridges (also evident in [7,25]) when they run along the prism–ipm boundaries as shown
in figure 8e. Unbroken prism ligaments (a few prisms), as demonstrated in figure 8f , are also
reported in various studies of human and bovine enamel and have a significant contribution to
the fracture toughness by shielding the crack tip and further reducing the local stress intensity
available to drive fracture [7,8,11,12,25]. Observations of the crack paths in transversal oriented
samples revealed crack branching (figure 8b) as a major visible toughening mechanism [8,11,12],
whereas the term ‘meandering cracks’ was introduced [11] to describe branches that rejoin the
main crack again after some micrometres of propagation. Crack branching yields an increase
in the crack surface with increased energy consumption, which is otherwise available for crack
advance [66]. Finally, microcracking about the main crack is detected in enamel [7,9,12,16,25] that
is located in the prism boundaries and ipm as demonstrated in figure 8c. The contribution of
microcracking in the vicinity of a crack tip creates dilatation of the structure and as a result reduces
the effective elastic modulus in the process zone, thereby lowering the crack tip stresses and the
effective stress intensity [7,66]. In terms of crack growth, microcracks contributing to the total
crack surface are generated which consequently increases the energy required [66].
Downloaded from http://rsta.royalsocietypublishing.org/ on May 27, 2015

(d) Hierarchical level 3: enamel types 13


(i) Structural features

rsta.royalsocietypublishing.org Phil. Trans. R. Soc. A 373: 20140130


.........................................................
The enamel types in prismatic enamel are defined by whether the prisms are parallel to each
other (radial enamel) or cross each other with different orientations (decussating enamel), as
demonstrated in figure 1. Radial enamel is characterized by prisms developing radially in
accordance to the DEJ and is considered the most basic prismatic enamel type. Modified radial
enamel and additional enamel types have developed over time to meet specific biomechanical
demands related to change in feeding habits and body size that gave rise to higher and multi-
directional stresses acting on enamel [57,65]. The derived enamel types such as decussating
enamel did not replace radial enamel in many species but arose in combination with it,
forming a complex schmelzmuster [65]. There is a huge variety in prism decussation patterns
among different species, but the most commonly observed specific mode is HSBs, where prisms
decussate in alternating layers. HSBs are described by the inclination and thickness of the bands.
The inclination of the bands can be readily seen in vertical sections of teeth and they are found
to be primitively perpendicular to the DEJ, and inclined upwards to the OES in more advanced
states. The thickness of the bands is usually in the range of 10–20 prisms. Regular bifurcation of
the HSBs is reported by Koenigswald et al. [67] and regarded as another important feature of the
bands.

(ii) Implications for fracture behaviour


The stress–strain behaviour cannot be analysed for level 3 in this review as isolated
characterization of radial enamel and HSBs has not yet been published. Their dimensions (approx.
200–600 µm in cross section) complicate both the macro- and micro-scale sample preparation.
However, preparation of compact tension (CT) specimens was possible for R-curve tests.
Therefore, only the fracture toughness and toughening mechanisms associated with radial enamel
of parallel aligned prisms and decussating enamel (HSBs with periodically decussating prism
bundles) can be compared. The fracture toughness of isolated radial and decussating enamel were
quantified by Yahyazadehfar et al. [12] using incremental crack growth measurements with inset
CT specimens of human enamel and the results revealed that both types undergo an increase in
crack growth resistance; in other words, they exhibit rising R-curve behaviour. Both the initiation
toughness Ko and the maximum toughness Kc (the magnitude of stress intensity at the onset of
unstable fracture) are found to be substantially higher in decussating enamel (see the electronic
supplementary material, table S2), which can be attributed to the potency of the toughening
mechanisms prevailing in different enamel types [12].
The representative crack growth paths in the radial and decussating enamel resulting from
the aforementioned study are presented in the electronic supplementary material. The most
prominent toughening mechanism in radial enamel was identified to be crack branching, which
often served as a precursor to the generation of unbroken ligaments of prism bundles each
comprising a relatively limited number of prisms (5–10) [12]. However, within decussating
enamel, crack bridging was more dominant than crack branching and consisted of larger
unbroken prism bundles that comprised between 10 and 50 individual prisms [12]. More details
regarding the superiority of enamel with prism decussation over radial enamel in stopping crack
propagation are considered in the electronic supplementary material.

(e) Hierarchical level 4: schmelzmuster


(i) Structural features
The schmelzmuster describes the spatial distribution of different enamel types in a tooth
according to the hierarchical terminology formulated by Koenigswald & Clemens [14]. As noted
in the Introduction, various enamel types exist in a single tooth. For example, in most mammals,
Downloaded from http://rsta.royalsocietypublishing.org/ on May 27, 2015

schmelzmuster consists of radial enamel (parallel arranged prisms) in the outer and decussation
14
enamel (crossing groups of prisms) in the inner enamel [15] (figure 1). The consequences of this
configuration are elaborated later in §2e(ii). The change in the enamel types causes some other

rsta.royalsocietypublishing.org Phil. Trans. R. Soc. A 373: 20140130


.........................................................
compositional and structural alterations from outer to inner enamel. For instance, the prism–ipm
arrangement changes from a honeycomb pattern in the outer enamel to a sandwich structure
in the inner enamel (specific to bovine enamel). Additionally, enamel prisms do not simply
run straight from the DEJ towards the outer surface but they locally bend, twist and change
their position [68]. In human enamel, as a function of the ameloblast cell movements, strongly
undulating prisms in the inner enamel are observed (this is less pronounced in the inner region
of bovine enamel), whereas prisms follow a straight course in the cervical (innermost) and outer
regions of enamel [68]. Finally, it is well documented that the protein content of enamel increases
from the outer to inner region as verified by He et al. [6] based upon Raman spectroscopic analysis.

(ii) Implications for fracture behaviour


The role of the spatial distribution of different enamel types (with regards to the corresponding
structural and compositional variations from outer to inner enamel) in the fracture behaviour of
the tissue is elaborated in this section. For this purpose, stress–strain and crack growth resistance
measurements performed on bulk enamel samples, which represent the hierarchical level 4 (and
containing all the subordinate levels), are presented. To the best of the authors’ knowledge, there
exists only a single study reporting the strength of bulk enamel as 11–42 MPa with a tensile testing
set-up [36] (as illustrated in the electronic supplementary material, figure S1). The limitations
of these studies are that the deformation and resultant stress–strain curves were not measured.
Only a recent dissertation provided greater insight into the deformation and fracture behaviour
of enamel using flexure specimens from bovine incisors owing to their larger size and thicker
enamel layer [9]. Bending experiments were performed using a custom-made three-point bending
device with two sample groups designated as transversal and longitudinal. The orientations of
the prisms within each sample group are illustrated by the schematic drawings in figure 9a.
In the case of transversal samples, the prisms were oriented perpendicular to the long axis of
the sample and both the outer and inner enamel located on the tensile side of the specimen.
The longitudinal samples were tested in such a direction that cracks grew from the outer to the
inner enamel. The mean strength, elastic modulus and strain to failure values of six samples for
both groups are listed in the electronic supplementary material. The fracture strength and elastic
modulus of transversal samples were found to be 95 ± 7 MPa and 56 ± 7 GPa, respectively. These
are significantly higher than the fracture strength (69 ± 20 MPa) and elastic modulus (35 ± 6 GPa)
obtained from the longitudinal samples. Bechtle [9] explained the superior mechanical properties
of transversal samples based on the fact that the inner enamel is involved in the fracture process
from the beginning so that some prisms and ipm layers are broken during the entire fracture
process. Comparing the strength and elastic modulus of bulk enamel with bulk HAP, the values
of the former are slightly lower, and this can be attributed to the presence of proteins in enamel.
However, there is a one order of magnitude reduction in the fracture strength from the crystallite
level of approximately 1000 MPa to the bulk enamel of approximately 100 MPa. The orientation
of the structural elements (crystallites, prisms) is believed to play a major role, which is highly
complex in bulk enamel, and controlled in the small-scale samples shown in figure 4. In the
latter, crystallites and prisms were uniaxially loaded along their long axis, leading to mineral
fracture; however, in the bulk samples, the majority of the prisms were loaded perpendicular
to their long axis, leading to prism separation at the weaker protein boundaries. The fracture
strength of micro-cantilever bending specimens, when loaded perpendicular to prism axes, was
reduced to approximately 350 MPa [49], supporting our argument on the role of orientation.
Moreover, according to Gao’s hierarchical model, each additional level of hierarchy adds more
protein to the composite structure [16], which can also account for the much lower strength of bulk
enamel. Finally, in the millimetre-sized bulk samples, it is difficult to avoid microcracks induced
by successive grinding and polishing procedures. Accordingly, the amount and size of the defects
Downloaded from http://rsta.royalsocietypublishing.org/ on May 27, 2015

hierarchical level 4—schmelzmuster


15
(a) 120 (b) direction transversal
transversal of crack

rsta.royalsocietypublishing.org Phil. Trans. R. Soc. A 373: 20140130


.........................................................
inner enamel
100 outer enamel propagation

80
stress (MPa)

60 longitudinal
100 mm
outer enamel
40 inner enamel
(c) direction of crack longitudinal
propagation

20 outer enamel
inner enamel
0
0 0.2 0.4 0.6 0.8 1.0 1.2 1.4
strain (%)
250 mm

Figure 9. Results of the three-point bending tests of bulk enamel: representative stress–strain curves of bovine enamel tested
in longitudinal and transversal orientations illustrate the quasi-ductile deformation behaviour of bulk enamel (a) (image from
Bechtle [9]). The simplified schematic drawings of the sample groups show the arrangement of prisms, where the outer
and inner enamel regions are marked by light and dark grey, respectively. The post-test microscopic observations of the
representative transverse (b) and longitudinal (c) (image from Bechtle [9]) samples display different crack profiles governed
by the prism orientations. The micrograph of the transversal sample (1 × 1 × 10 mm3 ) shown in (b) was obtained in an
independent in situ three-point bending study. (Online version in colour.)

are substantially higher in bulk samples, which can also lower the strength. Nevertheless, the
benefit of hierarchy at this level manifested itself in the damage-tolerant stress–strain behaviour
(figure 9a). Neither of the bulk sample groups showed a catastrophic fracture but failed in a
gradual manner resembling those of artificial fibre-reinforced composites [9]. There was a sudden
drop in the stress level at fracture initiation but further loading resulted in progressive damaging
of the samples. Both enamel sample groups remained macroscopically intact after testing. The
crack profiles varied with respect to the prism orientation. In the transversal sample, multiple
macroscopic cracks spreading over a large area were evident (figure 9b); however, there was a
primary crack that propagated straight through the outer enamel which became highly tortuous
within the inner enamel of the longitudinal sample (figure 9c).
In general, apparent fracture toughness values obtained by different methods corresponding
to the hierarchical level 1 and level 2 lie roughly between 0.5 and 1.5 MPa m0.5 . At hierarchical
level 3, radial and decussating enamel were analysed separately and both exhibited rising crack
growth resistance, whereas R-curve studies at hierarchical level 4 investigated the synergistic
effect of radial and decussating enamel and the role of their spatial locations. Bechtle et al. [10,11]
examined the role of prism orientation (transversal and longitudinal) and crack growth direction
(from outer to inner enamel and the reverse direction) using single edge notched (SENB) bulk
bovine enamel samples tested in three-point bending. All cracks grew at an oblique angle of
approximately 45◦ from the initial notch direction independent of sample orientation with the
crack plane always being perpendicular to the side surface (as illustrated with the red line on
the schematic drawing in figure 10). This introduced mixed-mode stress conditions at the crack
tip, mode I (tensile) and mode II (in-plane shear) stresses. The determined KIR ∗ and K∗ values
IIR
increased with increasing crack extension in all sample groups (asterisk indicates that this is
not a fracture resistance in the usual sense because of the lack of a known fracture criterion
for enamel under mixed-mode loading). The initiation and maximum fracture toughness values
obtained from the R-curves are listed in the electronic supplementary material, table S2. KIIR ∗
Downloaded from http://rsta.royalsocietypublishing.org/ on May 27, 2015

hierarchical level 4—schmelzmuster


16
5.0
K*IR longitudinal
4.5

rsta.royalsocietypublishing.org Phil. Trans. R. Soc. A 373: 20140130


crack growth from outer

.........................................................
crack
4.0 to inner enamel
K*IR (MPa*m1/2)

3.5 notch
K*IR transversal
3.0
2.5 crack
2.0 notch
1.5 K*IR longitudinal
crack growth from inner
1.0
to outer enamel crack
0.5 range of KIC
0 indentation values notch
0 100 200 300 400 500 600
∆a (mm)

Figure 10. Fracture toughness resistance curves of bovine enamel samples (1 × 1 × 10 mm3 ) from the SENB three-point
bending method in comparison with published indentation toughness values indicated by the shaded box (image from Bechtle
[9]).

values were always about one-quarter to one-third of the KIR ∗ values for all orientations tested

by the same amount of crack increment and therefore only mode I fracture toughness resistance
curves are shown in figure 10. Also included in the graph are the corresponding crack lengths
(approx. 50 µm), the indentation toughness values obtained from enamel that correspond to the
initiation toughness at crack growth KIR ∗ = 1 MPa m0.5 of the SENB specimens. K∗ values rise to
IR
0.5
4–4.5 MPa m after 300–500 µm crack extension in transversal and longitudinal (crack growth
from outer to inner enamel) samples. However, in the other longitudinal samples with crack
growth from inner to outer enamel, the fracture toughness increased only to 2.5–3.0 MPa m0.5 .
Similarly, R-curve measurements performed by CT specimens revealed (results listed in the
electronic supplementary material, table S2) that enamel exhibits an increase in the fracture
resistance in both directions, but cracks growing from inner to outer enamel were clearly less
stable [7,13]. The reasons for higher fracture resistance of cracks propagating from outer to inner
enamel are considered in the electronic supplementary material.
Bajaj & Arola [7] estimated the contributions of the dominant toughening mechanisms to
the fracture toughness of enamel using theoretical models. The authors did not specify the
contributions according to the size scales but only according to the mechanisms. Unbroken
mineral ligaments have the highest contribution to toughening of enamel with bridging stress
intensity between 0.2 and 0.6 MPa m0.5 , which is approximately 25% of the global toughness in
their study [7,13]. The reduction in the local stress intensity factor induced by bridging of organic
matrix was estimated to be 0.1 MPa m0.5 . The contribution associated with microcracking was
estimated to range from 0.38 to 0.49 MPa m0.5 .

3. Conclusion
The hierarchical microstructure of enamel and the research on its fracture behaviour were
reviewed by classifying the results from independent studies according to their corresponding
hierarchical levels. This provided an overview of the influence of hierarchical organization
on the fracture toughness, fracture strength and deformation behaviour quantitatively. The
contribution of the structural features to the mechanical results was specified for each hierarchical
level. Overall, these results indicate that there is a sharp decrease in fracture strength from
nano-scaled HAP crystallites to bulk enamel containing all levels of hierarchy. But, there is
Downloaded from http://rsta.royalsocietypublishing.org/ on May 27, 2015

also a transition from brittle to quasi-ductile deformation behaviour indicating the damage
17
tolerance of enamel acquired by increasing hierarchy. This is also supported by the fracture
toughness investigations showing R-curve behaviour measured at higher hierarchical levels.

rsta.royalsocietypublishing.org Phil. Trans. R. Soc. A 373: 20140130


.........................................................
Toughening mechanisms including microcracking, protein and mineral bridges, crack deflection
and branching are identified at almost all hierarchical levels. It is believed that they act
simultaneously on the crack, resulting in growth resistance of enamel, but the contribution of
these mechanisms increases with increasing number of levels owing to the enlarged size of the
structural elements. Apart from the structural mechanisms gained by hierarchical organization,
increasing number of levels facilitates higher protein content as suggested by Gao’s hierarchical
model, which is however not as yet directly quantifiable. Although the quantitative contribution
of crack bridges formed by proteins is found to be insignificant [7], their benefit is postulated
to be limited slippage that could reduce stresses without crack growth and recovery (closure) of
cracks [53,69,70]. If there were no proteins, then there would be greater damage accumulation
and break-up of the microcracked structure, especially under multiple contact loading, which
is not frequently observed. With age, as the protein breaks down and is not repaired, then
enamel tends to become more brittle [71]. Overall, enamel consists of nanoscale building blocks
and weak interfaces over various length scales and has acquired, with evolution and species-
dependent nutrition, a complex architecture. This enables enamel to arrest cracks and to channel
them along configurations which eventually impede their propagation [72]. This strategy is
designated as ‘crack trapping and locking’ by Mirkhalaf et al. [72] and has more recently been
implemented in glass structures. The deformability of the glass was improved by generating weak
interfaces—through three-dimensional laser engraving—that were infiltrated with polyurethane.
The complex nature of the microstructure and resultant mechanical response of enamel,
especially crack growth behaviour with increasing hierarchical level, remains an area of fertile
research. In this study, only quasi-static crack growth has been investigated, whereas biologically
enamel experiences complex repetitive stresses as individual teeth are loaded. In addition, the
oral environment, with its range of pH, presence of proteases that may assist with break-up
of the remnant protein coupled with the current widespread usage of bleaching procedures
that fragment these proteins, may have an influence on the resultant properties, especially of
the higher hierarchical levels. The clinical implications of these aspects along with the role of
various clinical conditions such as enamel caries, hypo-mineralization, fluorosis and others awaits
further investigation.
Acknowledgements. We gratefully acknowledge Sabine Bechtle, Hüseyin Özcoban, Siang Fung Ang, Lihong He
and Zonghan Xie, who were involved in the experimental works presented here. We also thank Daniel Laipple
and Uwe Lorenz for technical support with the FIB system. Finally, we express gratitude to the German
Research Foundation (grant no. SCHN 372/18-1) for financial support on some of the investigations described
in this manuscript.

References
1. Elliott JC. 1997 Structure, crystal chemistry and density of enamel apatites. Ciba Found. Symp.
205, 54–67. (doi:10.1002/9780470515303.ch5)
2. Rasmussen ST, Patchin RE, Scott DB, Heuer AH. 1976 Fracture properties of human enamel
and dentin. J. Dent. Res. 55, 154–164. (doi:10.1177/00220345760550010901)
3. Hassan R, Caputo AA, Bunshah RF. 1981 Fracture toughness of human enamel. J. Dent. Res.
60, 820–827. (doi:10.1177/00220345810600040901)
4. Rasmussen S, Patchin R. 1984 Fracture properties of human enamel and dentin in an aqueous
environment. J. Dent. Res. 63, 1362–1368. (doi:10.1177/00220345840630120501)
5. Chai H, Lee JJ, Constantino PJ, Lucas PW, Lawn BR. 2009 Remarkable resilience of teeth. Proc.
Natl Acad. Sci. USA 106, 7289–7293. (doi:10.1073/pnas.0902466106)
6. He LH, Yin ZH, van Vuuren LJ, Carter EA, Liang XW. 2013 A natural functionally graded
biocomposite coating: human enamel. Acta Biomater. 9, 6330–6337. (doi:10.1016/j.actbio.
2012.12.029)
Downloaded from http://rsta.royalsocietypublishing.org/ on May 27, 2015

7. Bajaj D, Arola DD. 2009 On the R-curve behavior of human tooth enamel. Biomaterials 30,
18
4037–4046. (doi:10.1016/j.biomaterials.2009.04.017)
8. Bajaj D, Arola DD. 2009 Role of prism decussation on fatigue crack growth and fracture of

rsta.royalsocietypublishing.org Phil. Trans. R. Soc. A 373: 20140130


.........................................................
human enamel. Acta Biomater. 5, 3045–3056. (doi:10.1016/j.actbio.2009.04.013)
9. Bechtle S. 2011 On the mechanical properties of dental enamel: a multi-scale approach. PhD
dissertation. München: Verl. Dr. Hut. See http://www.worldcat.org/oclc/751020333.
10. Bechtle S, Fett T, Rizzi G, Habelitz S, Schneider GA. 2010 Mixed-mode stress intensity factors
for kink cracks with finite kink length loaded in tension and bending: application to dentin
and enamel. J. Mech. Behav. Biomed. Mater. 3, 303–312. (doi:10.1016/j.jmbbm.2009.12.004)
11. Bechtle S, Habelitz S, Klocke A, Fett T, Schneider GA. 2010 The fracture behaviour of dental
enamel. Biomaterials 31, 375–384. (doi:10.1016/j.biomaterials.2009.09.050)
12. Yahyazadehfar M, Bajaj D, Arola DD. 2013 Hidden contributions of the enamel
rods on the fracture resistance of human teeth. Acta Biomater. 9, 4806–4814.
(doi:10.1016/j.actbio.2012.09.020)
13. Yahyazadehfar M, Ivancik J, Majd H, An B, Zhang D, Arola D. 2014 On the mechanics of
fatigue and fracture in teeth. Appl. Mech. Rev. 66, 030803. (doi:10.1115/1.4027431)
14. Koenigswald WV, Clemens WA. 1992 Levels of complexity in the microstructure of
mammalian enamel and their application in studies of systematics. Scan. Microsc. 6, 195–217.
15. Maas MC, Dumont ER. 1999 Built to last: the structure, function, and evolution of
primate dental enamel. Evol. Anthropol. 8, 133–152. (doi:10.1002/(SICI)1520-6505(1999)8:
4<133::AID-EVAN4>3.0.CO;2-F)
16. Bechtle S, Ozcoban H, Lilleodden ET, Huber N, Schreyer A, Swain MV, Schneider GA. 2012
Hierarchical flexural strength of enamel: transition from brittle to damage-tolerant behaviour.
J. R. Soc. Interface 9, 1265–1274. (doi:10.1098/rsif.2011.0498)
17. Eastoe JE. 1960 Organic matrix of tooth enamel. Nature 187, 411–412. (doi:10.1038/187411b0)
18. Corcia JT, Moody WE. 1974 Thermal analysis of human dental enamel. J. Dent. Res. 53,
571–580. (doi:10.1177/00220345740530030901)
19. Holcomb DW, Young RA. 1980 Thermal decomposition of human tooth enamel. Calcif. Tissue
Int. 31, 189–201. (doi:10.1007/BF02407181)
20. Holager J. 1970 Thermogravimetric examination of enamel and dentin. J. Dent. Res. 49,
546–548. (doi:10.1177/00220345700490031301)
21. Glick PL. 1979 Patterns of enamel maturation. J. Dent. Res. 58, 883–895. (doi:10.1177/
00220345790580024201)
22. Larmas MA, Häyrynen H, Lajunen LHJ. 1993 Thermogravimetric studies on sound and
carious human enamel and dentin as well as hydroxyapatite. Eur. J. Oral Sci. 101, 185–191.
(doi:10.1111/j.1600-0722.1993.tb01102.x)
23. Macena MS, Leite ML, Gouveia CL, Lima TAS, Athayde PAA, Sousa FB. 2014 A comparative
study on component volumes from outer to inner dental enamel in relation to enamel tufts.
Arch. Oral Biol. 59, 568–577. (doi:10.1016/j.archoralbio.2014.03.001)
24. Little MF, Casciani FS. 1966 The nature of water in sound human enamel: a preliminary study.
Arch. Oral Biol. 11, 565–571. (doi:10.1016/0003-9969(66)90222-6)
25. Ang SF, Schulz A, Pacher Fernandes R, Schneider GA. 2011 Sub-10-micrometer toughening
and crack tip toughness of dental enamel. J. Mech. Behav. Biomed. Mater. 4, 423–432.
(doi:10.1016/j.jmbbm.2010.12.003)
26. Bechtle S et al. 2012 A method to determine site-specific, anisotropic fracture toughness in
biological materials. Scr. Mater. 66, 515–518. (doi:10.1016/j.scriptamat.2011.12.027)
27. Gupta H, Zioupos P. 2008 Fracture of bone tissue: the ‘hows’ and the ‘whys’. Med. Eng. Phys.
30, 1209–1226. (doi:10.1016/j.medengphy.2008.09.007)
28. Zhang Z, Zhang YW, Gao H. 2010 On optimal hierarchy of load-bearing biological materials.
Proc. R. Soc. B 278, 519–525. (doi:10.1098/rspb.2010.1093)
29. Jäger I, Fratzl P. 2000 Mineralized collagen fibrils: a mechanical model with staggered
arrangement of mineral particles. Biophys. J. 79, 1737–1746. (doi:10.1016/S0006-3495(00)
76426-5)
30. Gao H, Ji B, Jager IL, Arzt E, Fratzl P. 2003 Materials become insensitive to flaws at nanoscale:
lessons from nature. Proc. Natl Acad. Sci. USA 100, 5597–5600. (doi:10.1073/pnas.0631609100)
31. Gao H. 2006 Application of fracture mechanics concepts to hierarchical biomechanics of bone
and bone-like materials. Int. J. Fract. 138, 101–137. (doi:10.1007/s10704-006-7156-4)
32. Ji B, Gao H. 2010 Mechanical principles of biological nanocomposites. Annu. Rev. Mater. 40,
77–100. (doi:10.1146/annurev-matsci-070909-104424)
Downloaded from http://rsta.royalsocietypublishing.org/ on May 27, 2015

33. Yao H, Gao H. 2007 Multi-scale cohesive laws in hierarchical materials. Int. J. Solids Struct. 44,
19
8177–8193. (doi:10.1016/j.ijsolstr.2007.06.007)
34. Xu H, Smith D, Jahanmir S, Romberg E, Kelly J, Thompson V, Rekow ED. 1998 Indentation

rsta.royalsocietypublishing.org Phil. Trans. R. Soc. A 373: 20140130


.........................................................
damage and mechanical properties of human enamel and dentin. J. Dent. Res. 77, 472–480.
(doi:10.1177/00220345980770030601)
35. White SN, Luo W, Paine ML, Fong H, Sarikaya M, Snead ML. 2001 Biological organization
of hydroxyapatite crystallites into a fibrous continuum toughens and controls anisotropy in
human enamel. J. Dent. Res. 80, 321–326. (doi:10.1177/00220345010800010501)
36. Giannini M, Soares CJ, Carvalho RM. 2004 Ultimate tensile strength of tooth structures. Dent.
Mater. 20, 322–329. (doi:10.1016/S0109-5641(03)00110-6)
37. Daculsi G, Kerebel B. 1978 High-resolution electron microscope study of human enamel
crystallites: size, shape, and growth. J. Ultrastruct. Res. 65, 163–172. (doi:10.1016/S0022-
5320(78)90053-9)
38. Daculsi G, Menanteau J, Kerebel LM, Mitre D. 1984 Length and shape of enamel crystals.
Calcif. Tissue Int. 36, 550–555. (doi:10.1007/BF02405364)
39. Nanci A. 2007 Enamel: composition, formation, and structure. In Ten Cates’ oral histology:
development, structure, and function, 7th edn. St. Louis, MO: Mosby.
40. Kerebel B, Daculsi G, Kerebel LM. 1979 Ultrastructural studies of enamel crystallites. J. Dent.
Res. 58, 844–851. (doi:10.1177/00220345790580023701)
41. Nylen M. 1979 Matrix-mineral relationships-A. Morphologist’s viewpoint. J. Dent. Res. 58,
922–929. (doi:10.1177/00220345790580024601)
42. Nakahira A, Tamai M, Miki S, Pezzotti G. 2002 Fracture behavior and biocompatibility
evaluation of nylon-infiltrated porous hydroxyapatite. J. Mater. Sci. 37, 4425–4430.
(doi:10.1013/A:1020681309572)
43. Viswanath B, Raghavan R, Ramamurty U, Ravishankar N. 2007 Mechanical properties and
anisotropy in hydroxyapatite single crystals. Scr. Mater. 57, 361–364. (doi:10.1016/j.scriptamat.
2007.04.027)
44. Akao M, Aoki H, Kato K. 1981 Mechanical properties of sintered hydroxyapatite for prosthetic
applications. J. Mater. Sci. 16, 809–812. (doi:10.1007/BF00552220)
45. Selvig KA, Halse A. 1972 Crystal growth in rat incisor enamel. Anat. Rec. 173, 453–68.
(doi:10.1002/ar.1091730406)
46. Frazier PD. 1968 Adult human enamel: an electron microscope study of crystallite size and
morphology. J. Ultrastruct. Res. 22, 1–11. (doi:10.1016/S0022-5320(68)90045-29)
47. Bai P, Warshawsky H. 1985 Morphological studies on the distribution of enamel matrix
proteins using routine electron microscopy and freeze-fracture replicas in the rat incisor. Anat.
Rec. 212, 1–16. (doi:10.1002/ar.1092120102)
48. Moradian-Oldak J. 2001 Amelogenins: assembly, processing and control of crystal
morphology. Matrix Biol. 20, 293–305. (doi:10.1016/S0945-053X(01)00154-8)
49. Chan Y, Ngan A, King N. 2009 Use of focused ion beam milling for investigating the
mechanical properties of biological tissues: a study of human primary molars. J. Mech. Behav.
Biomed. Mater. 2, 375–383. (doi:10.1016/j.jmbbm.2009.01.006)
50. Chan Y, Ngan A, King N. 2011 Nano-scale structure and mechanical properties of the human
dentin–enamel junction. J. Mech. Behav. Biomed. Mater. 4, 785–795. (doi:10.1016/j.jmbbm.
2010.09.003)
51. Han L, Wang L, Song J, Boyce MC, Ortiz C. 2011 Direct quantification of the mechanical
anisotropy and fracture of an individual exoskeleton layer via uniaxial compression of
micropillars. Nano Lett. 11, 3868–3874. (doi:10.1021/nl201968u)
52. Volkert CA, Buch S, Heiland B, Dehm G. 2004 Transmission electron microscopy of
fluorapatite-gelatine composite particles prepared using focused ion beam milling. J. Microsc.
214, 208–212. (doi:10.1111/j.0022-2720.2004.01352.x)
53. He LH, Swain MV. 2007 Influence of environment on the mechanical behaviour of mature
human enamel. Biomaterials 28, 4512–4520. (doi:10.1016/j.biomaterials.2007.06.020)
54. Hull D, Clyne TW. 1996 An introduction to composite materials, pp. 158–171. Cambridge, UK:
Cambridge University Press.
55. Chieruzzi M, Pagano S, Pennacchi M, Lombardo G, D’Errico P, Kenny JM. 2012
Compressive and flexural behaviour of fibre reinforced endodontic posts. J. Dent. 40, 968–978.
(doi:10.1016/j.jdent.2012.08.003)
56. Xie Z, Swain MV, Hoffman MJ. 2009 Structural integrity of enamel: experimental and
modeling. J. Dent. Res. 88, 529–533. (doi:10.1177/0022034509337130)
Downloaded from http://rsta.royalsocietypublishing.org/ on May 27, 2015

57. Rensberger JM. 2000 Pathways to functional differentiation in mammalian enamel. In


20
Development, function and evolution of teeth (eds MF Teaford, MM Smith, MWJ Ferguson), pp.
252–268. Cambridge, UK: Cambridge University Press.

rsta.royalsocietypublishing.org Phil. Trans. R. Soc. A 373: 20140130


.........................................................
58. Line SR, Novaes PD. 2005 The development and evolution of mammalian enamel: structure
and functional aspects. Braz. J. Morphol. Sci. 22, 67–72.
59. Kieser JA, He L, Dean MC, Jones ME, Duncan WJ, Swain MV, Nelson NJ. 2011 Structure and
compositional characteristics of caniniform dental enamel in the tuatara Sphenodon punctatus
(Lepidosauria: Rhynchocephalia). N.Z. Dent. J. 107, 44–50.
60. Xie Z, Mahoney E, Kilpatrick N, Swain M, Hoffman M. 2007 On the structure–
property relationship of sound and hypomineralized enamel. Acta Biomater. 3, 865–872.
(doi:10.1016/j.actbio.2007.05.007)
61. Dusevich V, Xu C, Wang Y, Walker MP, Gorski JP. 2012 Identification of a protein-containing
enamel matrix layer which bridges with the dentin–enamel junction of adult human teeth.
Arch. Oral Biol. 57, 1585–1594. (doi:10.1016/j.archoralbio.2012.04.014)
62. Hu C et al. 1997 Sheathlin: cloning, cDNA/polypeptide sequences, and immunolocalization of
porcine enamel sheath proteins. J. Dent. Res. 76, 648–657. (doi:10.1177/00220345970760020501)
63. Elfallah HM, Swain MV. 2013 A review of the effect of vital teeth bleaching on the mechanical
properties of tooth enamel. N.Z. Dent. J. 109, 87–96.
64. Boyde A, Martin L. 1984 The microstructure of primate dental enamel. In Food acquisition and
processing in primates (eds DJ Chivers, BA Wood, A Bilsborough), pp. 341–367. Boston, MA:
Springer.
65. Koenigswald WV. 2000 Two different strategies in enamel differentiation: Marsupialia versus
Eutheria. In Development, function and evolution of teeth (eds MF Teaford, MM Smith, MWJ
Ferguson), pp. 107–118. Cambridge, UK: Cambridge University Press.
66. Munz D, Fett T. 1999 Ceramics: mechanical properties, failure behaviour, materials selection, pp. 61–
65. Berlin, Germany: Springer.
67. Koenigswald WV, Holbrook LT, Rose KD. 2011 Diversity and evolution of Hunter–Schreger
band configuration in tooth enamel of Perissodactyl mammals. Acta Palaeontol. Pol. 56, 11–32.
(doi:10.4202/app.2010.0021)
68. Radlanski RJ, Renz H, Willersinn U, Cordis CA, Duschner H. 2001 Outline and arrangement
of enamel rods in human deciduous and permanent enamel. 3D-reconstructions obtained
from CLSM and SEM images based on serial ground sections. Eur. J. Oral Sci. 109, 409–414.
(doi:10.1034/j.1600-0722.2001.00149.x)
69. He LH. 2008 Mechanical behaviour of human enamel and the relationship to its
structural and compositional characteristics. PhD dissertation, University of Sydney,
Faculty of Dentistry, Australia. See http://ses.library.usyd.edu.au/bitstream/2123/3536/1/
LH-He-2008-Thesis.pdf.
70. Rivera C, Arola D, Ossa A. 2013 Indentation damage and crack repair in human enamel.
J. Mech. Behav. Biomed. Mater. 21, 178–184. (doi:10.1016/j.jmbbm.2013.02.020)
71. Park S, Quinn J, Romberg E, Arola D. 2008 On the brittleness of enamel and selected dental
materials. Dent. Mater. 24, 1477–1485. (doi:10.1016/j.dental.2008.03.007)
72. Mirkhalaf M, Dastjerdi AK, Barthelat F. 2014 Overcoming the brittleness of glass through bio-
inspiration and micro-architecture. Nat. Commun. 5, 3166. (doi:10.1038/ncomms4166)

You might also like