Sie sind auf Seite 1von 11

See

discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/315318944

Computational Modeling of Interfacial


Behaviors in Nanocomposite Materials

Article in International Journal of Solids and Structures March 2017


DOI: 10.1016/j.ijsolstr.2017.02.029

CITATIONS READS

0 53

3 authors:

Liqiang Lin Xiaodu Wang


University of Texas at San Antonio University of Texas at San Antonio
9 PUBLICATIONS 21 CITATIONS 126 PUBLICATIONS 2,920 CITATIONS

SEE PROFILE SEE PROFILE

Xiaowei Zeng
University of Texas at San Antonio
33 PUBLICATIONS 286 CITATIONS

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Numerical Investigation of Collective Cell Migration View project

Multiscale Modeling of Ultrastructural Origins of Bone Fragility View project

All content following this page was uploaded by Liqiang Lin on 27 April 2017.

The user has requested enhancement of the downloaded file.


International Journal of Solids and Structures 115116 (2017) 4352

Contents lists available at ScienceDirect

International Journal of Solids and Structures


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

Computational modeling of interfacial behaviors in nanocomposite


materials
Liqiang Lin, Xiaodu Wang, Xiaowei Zeng
Department of Mechanical Engineering, University of Texas at San Antonio, TX 78249, United States

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

Article history: Towards understanding the bulk material response in nanocomposites, an interfacial zone model was pro-
Received 19 June 2016 posed to dene a variety of material interface behaviors (e.g. brittle, ductile, rubber-like, elastic-perfectly
Revised 18 January 2017
plastic behavior etc.). It also has the capability to predict bulk material response though independently
Available online 16 March 2017
control of the interface properties (e.g. stiffness, strength, toughness). The mechanical response of gran-
Keywords: ular nanocomposite (i.e. nacre) was investigated through modeling the relatively soft organic interface
Material interface modeling as an interfacial zone among hard mineral tablets and simulation results were compared with experi-
Organic interface mental measurements of stress-strain curves in tension and compression tests. Through modeling varies
Nacre material interfaces, we found out that the bulk material response of granular nanocomposite was reg-
Biological nanocomposite ulated by the interfacial behaviors. This interfacial zone model provides a possible numerical tool for
Polycrystalline structure qualitatively understanding of structure-property relationships through material interface design.
2017 Elsevier Ltd. All rights reserved.

1. Introduction To investigate how the interfacial behaviors and properties will


determine the mechanical performance of nanocomposite materi-
Biological nanocomposites such as sh scale, bone and nacre als, various experimental tests were conducted, it was found out
are increasingly attracting attention from engineers and re- that the toughness and strength of bulk materials can be improved
searchers due to their remarkable mechanical performances through different mechanisms, e.g. the interface shear deforma-
(Vernerey et al., 2014). In this kind of mineralized nanocomposites, tion (Barthelat and Espinosa, 2007; Gupta et al., 2006; Mercer
the volume fraction of mineral generally is over 85vol% (Dastjerdi et al., 2006), interlocking of nano-asperities and micro-scale wavi-
et al., 2013; Ritchie, 2011; Wang and Gupta, 2011). In some ex- ness (Espinosa et al., 2011; Wang et al., 2001), mineral bridging
treme cases, such as tooth enamel or shell of Strombus gigas, (Smith et al., 1999; Song et al., 2003). In the meanwhile, differ-
the volume fraction of mineral is up to 99vol% (Kamat et al., ent theoretical models have been proposed to study contributions
20 0 0; Yahyazadehfar and Arola, 2015). However, these biological of interfacial interaction on the mechanical response of bulk ma-
nanocomposites exhibit outstanding strength and toughness with terials (Okumura, 2015). The shear lag model was borrowed to
high contents of brittle minerals due to their ingenious design study the transfer of stress through tablets and interfaces under
of microstructures. The universal microstructural feature of these the tensile loading (Gao et al., 2003; Jackson et al., 1988; Kotha
biological nanocomposites is that the brittle mineral tablets are et al., 2001). The elastic energy model was applied to study the
bonded by relatively soft interfaces (Barthelat and Rabiei, 2011; effects of interface stiffness and volume fraction on the stiffness
Fratzl and Weinkamer, 2007; Gao, 2006; Ni et al., 2015; Vernerey and toughness of bulk materials (Okumura and De Gennes, 2001).
et al., 2014). Thus, although the minerals are brittle, the soft in- In addition, various numerical models were applied to study the
terfaces could deect the crack propagation and provide addition role of interfaces in determining the mechanical properties of bulk
energy dissipation paths along the interfacial surfaces. Through materials. The bilinear cohesive zone model (Geubelle and Bay-
the inspiration of wisdom and ingenious design in the biological lor, 1998) was used to investigate the effects of mineral-collagen
nanocomposites, a number of articial super materials have been interface behavior on microdamage accumulation in lamellar bone
fabricated (Bonderer et al., 2008; Deville et al., 2006; Finnemore tissues (Luo et al., 2011). The exponential cohesive zone model
et al., 2012; Munch et al., 2008). (Van den Bosch et al., 2006; Xu and Needleman, 1994) was em-
ployed to determine the independent roles of enzymatic and non-
enzymatic cross-linking on the mechanical behavior of a mineral-

Corresponding author.
ized collagen bril (Siegmund et al., 2008). Despite broad studies
E-mail address: xiaowei.zeng@utsa.edu (X. Zeng). have been conducted to scrutinize the interface contributions on

http://dx.doi.org/10.1016/j.ijsolstr.2017.02.029
0020-7683/ 2017 Elsevier Ltd. All rights reserved.
44 L. Lin et al. / International Journal of Solids and Structures 115116 (2017) 4352

the high mechanical performance of nanocomposites, it still has by relatively soft interfaces (Fig. 1). Although the interface is rel-
not been fully elucidated how the interfacial behaviors and prop- atively weak and its volume fraction is very low, it has high im-
erties determine the mechanical response of bulk materials. From pact on the mechanical performance of bulk material. During the
experimental perspective, it is dicult to directly observe the con- interface debonding process, the interface could undergo normal
tributions of interfacial properties on the mechanical response of or shear deformation. To describe the interface deformation behav-
bulk material because the deformation process involves multiscale iors, an interfacial zone model was proposed to mimic various in-
underlying mechanisms in strengthening toughness and strength terfacial behaviors. The traction-separation relations in normal and
(Barthelat and Rabiei, 2011; Peterlik et al., 2006). On the other shear directions were used to govern the interfacial behaviors as
hand, there are currently no comprehensive interfacial zone mod- shown in Fig. 2, which include four deformation stages in normal
els to capture the different interfacial behaviors in determination traction-separation law: compressive contact stage (0 0 ), elas-
of the bulk material response. tic stage ( 0 dn ), damage stage ( dn fn ) and complete failure
The aim of current work is to develop an interfacial zone model stage ( f n + ). In the shear traction-separation law, it has the
to describe difference material interface behaviors to predict the similar deformation stages (elastic stage (0 dt ), damage stage
bulk material response through material interface design. In this ( dt ft ) and complete failure stage ( ft )) with symmetry
work, the characteristics and features of the proposed interfacial in both positive and negative directions. The interface toughness is
zone model are discussed rstly. Then, the mechanical behavior the area under the traction-separation curve and the normal and
of granular nanocomposite (i.e. nacre) was studied using the pro- shear interface toughness are respectively dened as:
posed interfacial zone model to model the interfacial behaviors  fn
of nanocomposite. The simulation results provide insights on how n = T d (1)
the interfacial behaviors and interfacial damage pattern control the 0
 ft
bulk material response.
t = T d (2)
0

2. Development of the interfacial zone model In the interfacial interaction, it is not necessary for normal and
shear interaction always to have mutual effects on each other.
The granular biological nanocomposites (e.g. nacre) have the For instance, the uncramping and unfolding of molecule in colla-
universal structure in which the hard aragonite crystals are bonded gen brils due to stretching might not involve shear interaction

Fig. 1. Schematic of granular nanocomposite: (a) macroscopic structure of nacre, (b) microscopic structure of nacre.

Fig. 2. Generalized traction-separation relations used to describe interfacial behaviors: (a) normal direction, (b) tangential direction.
L. Lin et al. / International Journal of Solids and Structures 115116 (2017) 4352 45

Fig. 3. Different interfacial behaviors with positive shape parameters (pn , pt ).

Fig. 4. Different interfacial behaviors with negative shape parameters (pn , pt ).

(Meyers et al., 2013). For this reason, we presented an uncoupled between zero to one, the current model demonstrates a convex
interfacial zone model in this study for interfacial modeling. Based softening shape. When the shape parameters are equal to one,
on Xu and Needlemans exponential cohesive traction-separation the damage behavior of current model represents a linear relation.
law (Xu and Needleman, 1994), an interfacial zone model was pro- When the shape parameters are greater than one, the damage be-
posed to model different interfacial behaviors. The proposed inter- havior has a concave shape. As a special case, when pn = 0 or pt = 0,
facial traction-separation laws take the following form: it represents elastic-perfectly plastic behavior in normal or in shear
 n 0   0
 direction (Fig. 3). In addition, when shape parameters (pn , pt ) are
exp (1 )exp n n dn
dn
c
0

pn dn 0 taking as negative values, it can be seen from Fig. 4 that the inter-
Tn = c f n n dn < n < f n (3) face shows a rubber-like behavior.

f n n In addition, the stiffness of the interface at the equilibrium po-
0 n f n sition is dened as:



  dTn 1 n 0
2
(1 ) dtt exp 2t2 0 |t | dt Kn = = c exp(1 ) exp
c exp


p dt d n dn 0 dn 0
Tt = f t |t | t (4)  


t
c |t | f t dt dt < |t | < f t n 0 n 0  c exp(1 )
exp = (5)
0 n f t (dn 0 )2 dn 0  = dn 0
n 0

where c and c are cohesive strength in normal and tangential 


dTt 1 2
direction, respectively; and dn and dt are critical separation. The Kt = = c exp(1 ) exp t2
fn and ft are normal and tangential failure separation, respec- dt dt 2dt

tively. The 0 is the equilibrium position and exp (1) = 2.71828. The t 2
t
2  c exp(1 )
exp  = (6)
shape parameters (pn , pt ) are introduced to describe different in-
dt 3 2dt
2  dt
terfacial behaviors (or different traction-separation curve shapes),  =0 t

e.g. brittle, ductile, rubber-like behavior etc. In the model, the vari- Eqs. (5) and (6) suggest that a smaller value of ( dn 0 ) or dt
ables n and t represent normal and tangential separation, re- would result in a higher interface stiffness in normal or tangential
spectively. direction. Thus, ( dn 0 ) and dt can be used to tune the interface
The interfacial traction-separation laws for different positive stiffness in the normal and tangential directions.
and negative value of shape parameters (pn , pt ) are plotted In the proposed interfacial zone model, there are ve indepen-
in Figs. 3 and 4. It can be seen from Fig. 3 that the pro- dent parameters ( c , dn , 0 , fn and pn ) in normal direction and
posed interfacial zone model has the exponential properties be- four independent parameters ( c , dt , ft and pt ) in tangential di-
fore damage initiation position( dn or dt ). In the damage stage rection to manipulate different interfacial properties and behaviors.
(dn f n or dt f t ), the model shows different shapes with dif- The interface toughness (n , t ) can be readily determined from
ferent shape parameters (pn , pt ). When the shape parameter is the traction-separation curves.
46 L. Lin et al. / International Journal of Solids and Structures 115116 (2017) 4352

Fig. 5. 2D geometry model of a granular nanocomposite. The interfacial zone was generated by recessing the outline of the original Voronoi cells (dash line) to form the
new outline of the mineral grain (solid line). The grains in specimen were meshed by triangular element.

To test the proposed interfacial zone model, a two-dimensional modulus E = 100 GPa, Poissons ratio v = 0.28, mineral density = 3,
geometry of granular nanocomposite was generated in which the 190 kg/m3 . The soft interface in nacre contains different molecules
mineral crystals are bounded by an interfacial zone. The proposed and a direct experimental measurement of interface properties and
traction-separation laws were employed to mimic different inter- behaviors at nanoscale is very challenging. In this study, the me-
facial behaviors in the granular nanocomposite. The proposed in- chanical behaviors of the interface were simplied and governed
terfacial model was veried through a comparing study with ex- by the traction-separation laws, which can be dened through
perimental measurements (i.e. stress-strain relation). On the other those independent parameters described earlier. Those indepen-
hand, by using this model, the bulk mechanical behavior of granu- dent interfacial parameters will be estimated based on the experi-
lar nanocomposite can be predicted through material interface de- mental data and numerical simulation results reported in the liter-
sign. atures for different interfacial properties of nacre (Barthelat et al.,
2007; Espinosa et al., 2009; Evans et al., 2001; Lin and Meyers,
3. Geometry model and materials 2009; Meyers et al., 2013; Smith et al., 1999; Wang et al., 2001).

3.1. Geometry model of granular nanocomposite

4. Numerical simulations
A well-known example of granular nanocomposites is nacre.
The nacre is composed of aragonite (CaCO3 ) crystals, which are
organized in polygonal tablet, arranged in a thin layer of inter- 4.1. FEM implementations
faces (Fig. 5). For simplicity while ensuring reasonable accuracy, a
Following standard procedures and neglecting the body force,
two-dimensional (2D) model of CaCO3 crystals bounded through
a thin interface (Fig. 5) was generated in this study to represent the principle of virtual work of nite element formulation is writ-
ten as:
the microstructure of nacre. Briey, polygonal shaped CaCO3 crys-
tals were rst generated using the Centroidal Voronoi tessella-    
tion method in a 2D geometric model (Lin et al., 2014b). Then, P : Fd  Tinter dS = T udS u ud
an interfacial zone was created to represent the organic inter-  inter ext 
face between the CaCO3 crystals by recessing the edges of each (7)
grain cell in parallel towards the centroid of the grains with a
designated distance (Fig. 5). Based on the previous study regard- where ,  inter ,  ext are the volume, interface boundary and ex-
ing the minimum representative elementary volume (Evesque and ternal traction boundary of element in the reference conguration,
Adjmian, 20 02), 10 0 grains were required for the proposed 2D respectively; P is the rst Piola-Kirchhoff stress tensor; F is the de-
model to ensure consistent outcomes. In this study, roughly one formation gradient;  denotes the interfacial displacement jump
hundred and forty four (144) CaCO3 grains were generated in each across the interfaces; T denotes the external traction vector and
representative model. Based on the existing information on the Tinter is the interfacial bonding traction vector; represents the
thickness (20nm35 nm) (Barthelat and Espinosa, 2007; Barthelat material density in the reference conguration. The Newmark
et al., 2006; Dastjerdi et al., 2013; Lopez et al., 2014) and vol- method was applied for the explicit time integration with = 0
ume fraction (5% vol) (Barthelat et al., 2007; Song et al., 2003) and = 0.5 (Hughes et al., 1979).
of the organic interface in nacre, the interfacial zone thickness in In this study, the organic interface behaviors were governed by
the model was set to be 20 nm and the specimen size was set to the traction-separation laws described earlier. The CaCO3 grains
be Lx Ly = 8400 nm 8400 nm. were meshed using triangular elements. All simulations were im-
plemented using a custom-made nite element package, which
3.2. Material properties had been developed by Zeng and his co-workers (Li et al., 2012; Lin
et al., 2014a, 2015; Lin and Zeng, 2015; Zeng and Li, 2010, 2012).
According to the material properties of aragonite crystal re- Uniaxial load (tensile or compressive loading) was applied to the
ported in the literatures (Askarinejad and Rahbar, 2015; Barthe- nite element model by assigning a uniform displacement on the
lat et al., 2006; Hrabnkov et al., 2013; Zhang and Chen, top and bottom edges of the model (Lin et al., 2014a), with the
2013), the material properties of CaCO3 were set to be: Youngs right and left side of the specimen being set free.
L. Lin et al. / International Journal of Solids and Structures 115116 (2017) 4352 47

Table 1.
Parameters of interface property with
elastic-perfectly plastic behavior.

n (MPa) 110
( dn 0 )(nm) 0.6
( fn 0 )(nm) 30
pn 0.0
t (MPa) 40
dt (nm) 0.6
ft (nm) 30
pt 0.0

4.2. Comparison study of the interfacial zone model via an


elastic-perfectly plastic behavior Fig. 7. The stress-strain relations in tension and compression tests: the sim-tension
represented simulation result in tension; the sim-compression represented simula-
tion result in compression; the exp-tension denoted experimental result in tension;
The organic interface in the nacre is a thin layer of biopolymer, the exp-compression denoted experimental result in compression. The experimental
which has the capability of very large extension through sequen- results in abalone nacre were reported in Wang et al. (2001).
tial uncramping and unfolding of modules (Mohanty et al., 2008;
Smith et al., 1999). It was noted that experimental data did not
show a signicant hardening over the large stretching test. In addi- In current study, grains in the specimen were meshed by tri-
tion, the interface provided the slipping possibility between miner- angular element (Fig. 5). Based on the previous convergence test
als (Wang et al., 2001). Some researchers described the interface to of similar models (Lin et al., 2017, 2016), the mesh size in current
have a long plateau after yield point (Barthelat et al., 2007; Evans work is around 150 nm and the time step is t = 1 10 10 s. We
et al., 2001). Hence, we assumed that the interface has an elastic- have plotted the stress-strain curves after calculations, the stress-
perfectly plastic behavior. Specically, the normal critical displace- strain curves in the early post yield state agreed well with ex-
ment ( dn 0 ) at the interface was set to be 0.6 nm. This value perimental measurements as shown in Fig. 7. The average stress
was estimated by assuming that the tensile yield stain of the in- was compiled as mean standard deviation in which 6 samples
terface was 3.0% based on the information reported in literatures with random polycrystalline grain distribution are considered in
(Barthelat et al., 2007; Dastjerdi et al., 2013). The data reported the calculation. The bulk stress showed an initial strain harden-
by Smith et al. (1999) illustrated that some molecules with large ing after yield point in tension (Fig. 7). However, in compression,
stretching capability is present in the interface. According to the the stress-strain curve showed a relatively linear pattern until fail-
study of (Mohanty et al., 2008; Smith et al., 1999), the effective ure. The average yield stress of approximately 100 MPa in tension
displacement of plateau region is from 30 nm to 100 nm. Through is very close to the experimental yield stress of 105 MPa. The ulti-
parametric study and sensitivity test, it was found that the simula- mate stress was approximately 452 93 MPa in compression, and
tion results matched with experimental data when we set normal the experimental value of 370 MPa when testing was discontinued
failure separation ( fn 0 ) = 30 nm. Hence, the normal failure sep- lies in the range. In addition, the stiffness obtained from the sim-
aration ( fn 0 ) at the interface was estimated to be 30 nm in this ulation matched with the experimental measurement value at ap-
study by assuming that the tensile failure strain of the interface proximately 70 GPa. The agreement between the simulation results
was approximately 150%. In the absence of experimental data on and experimental data suggested that the proposed interfacial zone
the interfacial shear behavior, we assumed the traction-separation model has the capability to capture the characteristics and features
relation in tangential direction is similar to the one in normal di- of the interfacial behaviors.
rection. The critical shear separation dt was set to be 0.6 nm and
the failure shear separation ft was set to be 30 nm. The strength 4.3. Prediction of bulk material response via a rubber-like interfacial
in normal direction was set to be c = 110 MPa and the shear inter- behavior model
face strength was set to be c = 40 MPa based on the experimental
data reported in literatures(Barthelat and Espinosa, 2007; Espinosa As we know, different biopolymer may show different mate-
et al., 2009; Evans et al., 2001; Lin and Meyers, 2009). The details rial behaviors. For instance, the spider silk shows a rubber-like be-
of the interface parameters were listed in Table 1 and the curves havior (Meyers et al., 2013). This kind of biopolymer also showed
of the traction-separation laws were plotted in Fig. 6. high plastic deformation. To understand the effects of interface

Fig. 6. The traction-separation relations with respect to elastic-perfectly plastic interfacial behavior: (a) normal direction, and (b) shear direction.
48 L. Lin et al. / International Journal of Solids and Structures 115116 (2017) 4352

Table 2. organic interface between the minerals inside the whelk eggs has
Parameters of interface property with
an overall rubber-like property. In compression test, however, the
rubber-like behavior.
stress-strain curve showed a relatively linear response until failure
n (MPa) 130 (Fig. 9(b)).
( dn 0 )(nm) 3
( fn 0 )(nm) 15
pn 1.5 4.4. Deformation & failure pattern from different loading conditions
t (MPa) 40
dt (nm) 3
It has been investigated that the loading mode dependence of
ft (nm) 15
pt 1.5 nacre mechanical behavior is manifested (Evans et al., 2001; Wang
and Gupta, 2011; Wang et al., 2001). The ultimate stress of nacre in
tension is smaller than the one in compress (Jacobs, 1990). On the
other hand, the stress-strain curve of nacre shows highly nonlin-
behavior on the mechanical response of biological nanocompos- ear behavior in tension(Barthelat and Espinosa, 2007; Evans et al.,
ites, we designed a case by assuming that the interface pos- 2001; Wang et al., 2001), whereas it shows linear response in com-
sessed rubber-like behavior in a biological nanocomposite. The in- pression (Barthelat et al., 2006; Evans et al., 2001; Wang et al.,
terface parameters were estimated based on the information re- 2001). Furthermore, the dilatation band was formed among in-
ported in literatures (Lin and Meyers, 2009; Meyers et al., 2013). tertablets when nacre is in tension (Wang et al., 2001). This is
The designed interface properties were set to be: c = 130 MPa, highly indicated that the fracture mechanism of nacre in tension
c = 40 MPa, ( dn 0 ) = 3 nm, dt = 3 nm at 15% yield point and vs. compression.
( fn 0 ) = 15 nm, ft = 15 nm at 75% failure position, pn = pt = 1.5 Through analysis of the fracture pattern in the specimen, we
based on the properties reported in the literatures (Lin and Mey- found that the cracks were initiated at the interface perpendicu-
ers, 2009; Meyers et al., 2013). The details of the interface pa- lar to the loading direction when the model was loaded in tension
rameters were listed in Table 2 and the curves of the traction- (Figs. 10 and 11). When the damage at the interfaces started to
separation laws were plotted in Fig. 8. According to the data re- initiate, the stress 22 released within the damage interface zone
ported by Meyers et al. (2013), the failure strain of interface with region. As the loading kept increasing, cracks nucleated and coa-
rubber-like behavior in this case was set as 0.63. From the traction- lesced, and nally traveled transversely passing through the spec-
separation curve, we can readily determine the ultimate stress of imen (Figs. 10 and 11). It seems that there is a tendency for the
interface with rubber-like behavior to be 1.4 GPa. crack initiation, nucleation and coalescence mainly in mode I. This
In observation of the stress-strain curves for tension and com- phenomenon is analogous to the observation of dilatation bands
pression tests, it was found that the bulk stress with rubber- in the nacre under tension test (Wang et al., 2001).
like interface also showed a rubber-like behavior in stress-strain When the specimen loaded in compression, intergranular cracks
relation under tensile loading (Fig. 9(a)). This result may ex- propagated along an inclined angle via relative sliding between
plain that the stress-strain relation of whelk eggs always presents the mineral crystals (Figs. 12 and 13). This deformation/failure ten-
a trend of rubber-like behavior under tension test in different dency of inclined crack propagation within interfaces allowed more
temperature condition (Meyers et al., 2013) possibly because the energy dissipation during the damage process.

Fig. 8. The traction-separation relations with respect to rubber-like interface behavior: (a) normal direction, and (b) shear direction.

Fig. 9. The stress-strain relation in tension and compression for rubber-like interface: (a) tension; (b) compression.
L. Lin et al. / International Journal of Solids and Structures 115116 (2017) 4352 49

Fig. 10. Snapshots of stress distribution( 22 ) with elastic-perfectly plastic interface under tensile loading: (a) b = 0.19%, (b) b = 1.02% (crack initiation), (c) b = 1.05% (crack
nucleation), (d) b = 1.1% (crack coalescence).

Fig. 11. Snapshots of stress distribution( 22 ) with rubber-like interface under tensile loading: (a) b = 2.07%, (b) b = 2.26% (crack initiation), (c) b = 2.34% (crack nucleation),
(d) b = 2.38% (crack coalescence).
50 L. Lin et al. / International Journal of Solids and Structures 115116 (2017) 4352

Fig. 12. Snapshots of stress distribution( 22 ) with elastic-perfectly plastic interface under compressive loading: (a) b = 0.78%, (b) b = 0.83% (crack initiation), (c) b = 0.85%,
(d) b = 0.88% (crack propagation along an incline angle within interface) (e) b = 0.92%, (f) b = 0.99%.

Fig. 13. Snapshots of stress distribution( 22 ) with rubber-like interface under compressive loading: (a) b = 2.34%, (b) b = 2.49% (crack initiation), (c) b = 2.55% (crack prop-
agation along an incline angle within interface), (d) b = 2.57%, (e) b = 2.58%, (f) b = 2.61%.

4.5. Bulk material response vs. different interfacial behaviors haviors in tension (Fig. 15), suggesting that the interface behavior
directly dictates the bulk material response.
To further prove that interface behavior directly inuences the
bulk material response, we conducted three simulations by assum-
ing that the interface possessed a brittle, elastic-perfectly plastic 5. Discussion and conclusions
and rubber-like behavior in a biological nanocomposite, as shown
in Fig. 14. The details of the interface parameters were listed in In this study, we have reported an interfacial zone model for
Table 3. Through these studies, we explored the contribution of in- interface modeling in granular nanocomposites to study the bulk
terfacial properties on the bulk material response. It is found that material response. The proposed interfacial model has the capabil-
the bulk stress-strain curves are very similar to the interfacial be- ity to describe different interfacial material behaviors, e.g. brittle,
L. Lin et al. / International Journal of Solids and Structures 115116 (2017) 4352 51

Fig. 14. The traction-separation relations with respect to different interfacial behaviors: (a) brittle, (b) elastic-perfectly plastic, (c) rubber-like.

Table 3. direction. Thus the interface behavior dominated the bulk material
Parameters of interface property with different behaviors.
response. However, in compression, the interface could provide
brittle elastic perfectly plastic rubber-like facility for the CaCO3 crystals to continually slide along the inter-
n (MPa) 130 130 130 face till shear damage. Under compressive loading, the interfaces
( dn 0 )(nm) 3 3 3 are mainly in compression. Due to the relatively linear behavior of
( fn 0 )(nm) 3 30 15 interfaces under compression in normal direction, it might be the
pn 0 1.5 reason that the stress-strain curves showed a linear relation until
t (MPa) 40 40 40
failure in compression. Also, CaCO3 is a linear elastic material,
dt (nm) 3 3 3
ft (nm) 3 30 15 which may have contributions on the linear material response.
pt 0 1.5 There are several limitations associated with current work.
Firstly, a 2D plane strain model is used in the study of nacre, which
may not be fully representative of 3D cases. Although 3D mod-
els composed of multiple layers, the inner plane fracture between
ductile, elastic-perfectly plastic and rubber-like behaviors through mineral tablets is still the dominant failure mode in the struc-
adjusting the control parameters. ture based on experimental observations (Evans et al., 2001; Wang
Via comparison with experimental data, the stress-strain curves et al., 2001). In the early post yield state, the sliding between inter-
of simulation in the early post yield state agreed well with experi- layers might have minor effects on the bulk mechanical response,
mental measurements (Fig. 7). From the simulation results, we ob- thus the current 2D plane strain model is presumably only su-
served that the bulk stress-strain relation represented a rubber-like cient in capturing the early stage of post yield behaviors. Secondly,
response with rubber-like interface behavior in tension, but bulk the organic interface properties are simply estimated based on the
stress-strain showed a linear behavior in compression (Fig. 9). The experimental observations and related information reported in the
simulation results also captured the damage process in a 2D nacre literatures, which do not consider complex chemical bonding and
specimen. The loading mode dependence of the mechanical behav- environment of organic-inorganic interface and may be used only
ior of granular nanocomposites is manifested in the distinct defor- for qualitative analysis. Nonetheless, the results of this study indi-
mation mode of the organic interface; that is, intergranular open- cate that the proposed interfacial zone model is still able to cap-
ing mode in tension and intergranular sliding mode in compression ture the major deformation behavior of nacre, which may give rise
(Figs. 1013). to the insights on the mechanisms of superior tough and strong
From fracture pattern (Figs. 1013) and the stress-strain relation mechanical properties of nacre.
analysis (Figs. 7 and 9), we observed that the bulk material behav- The proposed interfacial zone model provides a numerical
ior was regulated by the interface properties. The possible reason tool to study bulk material response through dening different
is that the interface carried the most loads under tensile loading material interface behaviors. In addition, by adjusting the con-
due to its weak strength, which is easier to transfer loads to trollable parameters in the current interfacial zone model, we
adjacent interfaces after it damaged along the interface in normal could independently control the interface stiffness, toughness, and

Fig. 15. The stress-strain relation of bulk material with different interfacial properties: (a) interface with brittle behavior; (b) interface with elastic perfectly plastic behavior;
(c) interface with rubber-like behavior.
52 L. Lin et al. / International Journal of Solids and Structures 115116 (2017) 4352

mechanical strength in normal and shear directions. Through dif- Kotha, S., Li, Y., Guzelsu, N., 2001. Micromechanical model of nacre tested in tension.
ferent interfacial behavior modeling by the proposed model, it is J. Mater. Sci. 36.
Li, S., Zeng, X., Ren, B., Qian, J., Zhang, J., Jha, A.K., 2012. An atomistic-based inter-
indicated that the bulk material response is largely determined by phase zone model for crystalline solids. Comput. Meth. Appl. Mech. Eng. 229,
interfacial behavior. Through the model and simulation study, we 87109.
could predict the bulk material behavior by interfacial property. Lin, A.Y.-M., Meyers, M.A., 2009. Interfacial shear strength in abalone nacre. J. Mech.
Behav. Biomed. Mater. 2, 607612.
The main point of this paper is to demonstrate the capability of Lin, L., Dhanawade, R., Zeng, X., 2014a. Numerical simulations of dynamic fracture
the interfacial zone model and we believe the model can be used growth based on a cohesive zone model with microcracks. J. Nanomech. Mi-
to design articial tunable nanocomposite materials through mate- cromech. 4 B4014003.
Lin, L., Samuel, J., Zeng, X., Wang, X., 2017. Contribution of extrabrillar matrix to
rial interface design.
the mechanical behavior of bone using a novel cohesive nite element model.
J. Mech. Behav. Biomed. Mater. 65, 224235.
Acknowledgements Lin, L., Wang, X., Zeng, X., 2014b. Geometrical modeling of cell division and cell
remodeling based on Voronoi tessellation method. CMES 98, 203220.
Lin, L., Wang, X., Zeng, X., 2015. The role of cohesive zone properties on intergran-
This work is partially supported by a grant from National In- ular to transgranular fracture transition in polycrystalline solids. Int. J. Damage
stitutes of Health (Grant No. R21AR066925), a grant from National Mech. 1056789515618732.
Science Foundation (Grant No. CMMI-1538448), and a grant from Lin, L., Wang, X., Zeng, X., 2016. An improved interfacial bonding model for material
interface modeling. Eng. Fract. Mech..
the University of Texas at San Antonio, Oce of the Vice President Lin, L., Zeng, X., 2015. Computational modeling and simulation of spall fracture in
for Research. polycrystalline solids by an atomistic-based interfacial zone model. Eng. Fract.
Mech. 142, 5063.
References Lopez, M.I., Martinez, P.E.M., Meyers, M.A., 2014. Organic interlamellar layers, meso-
layers and mineral nanobridges: Contribution to strength in abalone (Haliotis
Askarinejad, S., Rahbar, N., 2015. Toughening mechanisms in bioinspired multilay- rufescence) nacre. Acta Biomater. 10, 20562064.
ered materials. J. R. Soc. Interf. 12, 20140855. Luo, Q., Nakade, R., Dong, X., Rong, Q., Wang, X., 2011. Effect of mineralcollagen in-
Barthelat, F., Espinosa, H., 2007. An experimental investigation of deformation and terfacial behavior on the microdamage progression in bone using a probabilistic
fracture of nacremother of pearl. Exp. Mech. 47, 311324. cohesive nite element model. J. Mech. Behav. Biomed. Mater. 4, 943952.
Barthelat, F., Li, C.-M., Comi, C., Espinosa, H.D., 2006. Mechanical properties of nacre Mercer, C., He, M., Wang, R., Evans, A., 2006. Mechanisms governing the inelastic
constituents and their impact on mechanical performance. J. Mater. Res. 21, deformation of cortical bone and application to trabecular bone. Acta Biomater.
19771986. 2, 5968.
Barthelat, F., Rabiei, R., 2011. Toughness amplication in natural composites. J. Mech. Meyers, M.A., McKittrick, J., Chen, P.-Y., 2013. Structural biological materials: critical
Phys. Solids 59, 829840. mechanics-materials connections. Science 339, 773779.
Barthelat, F., Tang, H., Zavattieri, P., Li, C.-M., Espinosa, H., 2007. On the mechanics Mohanty, B., Katti, K.S., Katti, D.R., 2008. Experimental investigation of nanomechan-
of mother-of-pearl: a key feature in the material hierarchical structure. J. Mech. ics of the mineral-protein interface in nacre. Mech. Res. Commun. 35, 1723.
Phys. Solids 55, 306337. Munch, E., Launey, M.E., Alsem, D.H., Saiz, E., Tomsia, A.P., Ritchie, R.O., 2008. Tough,
Bonderer, L.J., Studart, A.R., Gauckler, L.J., 2008. Bioinspired design and assembly of bio-inspired hybrid materials. Science 322, 15161520.
platelet reinforced polymer lms. Science 319, 10691073. Ni, Y., Song, Z., Jiang, H., Yu, S.-H., He, L., 2015. Optimization design of strong and
Dastjerdi, A.K., Rabiei, R., Barthelat, F., 2013. The weak interfaces within tough nat- tough nacreous nanocomposites through tuning characteristic lengths. J. Mech.
ural composites: experiments on three types of nacre. J. Mech. Behav. Biomed. Phys. Solids 81, 4157.
Mater. 19, 5060. Okumura, K., 2015. Strength and toughness of biocomposites consisting of soft and
Deville, S., Saiz, E., Nalla, R.K., Tomsia, A.P., 2006. Freezing as a path to build com- hard elements: a few fundamental models. MRS Bull. 40, 333339.
plex composites. Science 311, 515518. Okumura, K., De Gennes, P.-G., 2001. Why is nacre strong? Elastic theory and frac-
Espinosa, H.D., Juster, A.L., Latourte, F.J., Loh, O.Y., Gregoire, D., Zavattieri, P.D., 2011. ture mechanics for biocomposites with stratied structures. Eur. Phys. J. E 4,
Tablet-level origin of toughening in abalone shells and translation to synthetic 121127.
composite materials. Nat. commun. 2, 173. Peterlik, H., Roschger, P., Klaushofer, K., Fratzl, P., 2006. From brittle to ductile frac-
Espinosa, H.D., Rim, J.E., Barthelat, F., Buehler, M.J., 2009. Merger of structure and ture of bone. Nat. Mater. 5, 5255.
material in nacre and bonePerspectives on de novo biomimetic materials. Prog. Ritchie, R.O., 2011. The conicts between strength and toughness. Nat. Mater. 10,
Mater Sci. 54, 10591100. 817822.
Evans, A., Suo, Z., Wang, R., Aksay, I., He, M., Hutchinson, J., 2001. Model for the Siegmund, T., Allen, M.R., Burr, D.B., 2008. Failure of mineralized collagen brils:
robust mechanical behavior of nacre. J. Mater. Res. 16, 24752484. modeling the role of collagen cross-linking. J. Biomech. 41, 14271435.
Evesque, P., Adjmian, F., 2002. Stress uctuations and macroscopic stick-slip in Smith, B.L., Schffer, T.E., Viani, M., Thompson, J.B., Frederick, N.A., Kindt, J.,
granular materials. Eur. Phys. J. E 9, 253259. Belcher, A., Stucky, G.D., Morse, D.E., Hansma, P.K., 1999. Molecular mechanis-
Finnemore, A., Cunha, P., Shean, T., Vignolini, S., Guldin, S., Oyen, M., Steiner, U., tic origin of the toughness of natural adhesives, bres and composites. Nature
2012. Biomimetic layer-by-layer assembly of articial nacre. Nat. commun. 3, 399, 761763.
966. Song, F., Soh, A., Bai, Y., 2003. Structural and mechanical properties of the organic
Fratzl, P., Weinkamer, R., 2007. Natures hierarchical materials. Prog. Mater Sci. 52, matrix layers of nacre. Biomaterials. 24, 36233631.
12631334. Van den Bosch, M, Schreurs, P., Geers, M., 2006. An improved description of the
Gao, H., 2006. Application of fracture mechanics concepts to hierarchical biome- exponential Xu and Needleman cohesive zone law for mixed-mode decohesion.
chanics of bone and bone-like materials. Int. J. Fract. 138, 101137. Eng. Fract. Mech. 73, 12201234.
Gao, H., Ji, B., Jger, I.L., Arzt, E., Fratzl, P., 2003. Materials become insensitive to Vernerey, F.J., Musiket, K., Barthelat, F., 2014. Mechanics of sh skin: a compu-
aws at nanoscale: lessons from nature. Proc. Natl. Acad. Sci. 100, 55975600. tational approach for bio-inspired exible composites. Int. J. Solids Struct. 51,
Geubelle, P.H., Baylor, J.S., 1998. Impact-induced delamination of composites: a 2D 274283.
simulation. Composites Part B 29, 589602. Wang, R., Gupta, H.S., 2011. Deformation and fracture mechanisms of bone and
Gupta, H.S., Seto, J., Wagermaier, W., Zaslansky, P., Boesecke, P., Fratzl, P., 2006. Co- nacre. Annu. Rev. Mater. Res. 41, 4173.
operative deformation of mineral and collagen in bone at the nanoscale. Proc. Wang, R., Suo, Z., Evans, A., Yao, N., Aksay, I., 2001. Deformation mechanisms in
Natl. Acad. Sci. 103, 1774117746. nacre. J. Mater. Res. 16, 24852493.
Hrabnkov, I., Frda, J., epitka, J., Sasaki, T., Frdov, B., Luke, J., 2013. Mechani- Xu, X.-P., Needleman, A., 1994. Numerical simulations of fast crack growth in brittle
cal properties of deep-sea molluscan shell. Comput. Methods Biomech. Biomed. solids. J. Mech. Phys. Solids 42, 13971434.
Engin. 16, 287289. Yahyazadehfar, M., Arola, D., 2015. The role of organic proteins on the crack growth
Hughes, T.J., Pister, K.S., Taylor, R.L., 1979. Implicit-explicit nite elements in nonlin- resistance of human enamel. Acta Biomater. 19, 3345.
ear transient analysis. Comput. Meth. Appl. Mech. Eng. 17, 159182. Zeng, X., Li, S., 2010. A multiscale cohesive zone model and simulations of fractures.
Jackson, A., Vincent, J., Turner, R., 1988. The mechanical design of nacre. Proc. R. Comput. Meth. Appl. Mech. Eng. 199, 547556.
Soc. Lond. B 234, 415440. Zeng, X., Li, S., 2012. Application of a multiscale cohesive zone method to model
Jacobs, D.K., 1990. Sutural pattern and shell stress in Baculites with implications for composite materials. Int. J. Multiscale Comput. Eng. 10, 391405.
other cephalopod shell morphologies. Paleobiology 16, 336348. Zhang, N., Chen, Y., 2013. Nanoscale plastic deformation mechanism in single crystal
Kamat, S., Su, X., Ballarini, R., Heuer, A., 20 0 0. Structural basis for the fracture aragonite. J. Mater. Sci. 48, 785796.
toughness of the shell of the conch Strombus gigas. Nature 405, 10361040.

View publication stats

Das könnte Ihnen auch gefallen