7
Mechanical properties of a hollow-cylindrical-joint honeycomb Qiang Chen a , Nicola Pugno b,c,d , Kai Zhao e , Zhiyong Li a,a Biomechanics Laboratory, School of Biological Science and Medical Engineering, Southeast University, 210096 Nanjing, PR China b Laboratory of Bio-Inspired & Graphene Nanomechanics, Department of Civil, Environmental and Mechanical Engineering, University of Trento, I-38123 Trento, Italy c Center for Materials and Microsystems, Fondazione Bruno Kessler, I-38123 Trento, Italy d School of Engineering and Materials Science, Queen Mary Univerisity of London, Mile End Road, London, E1 4NS, UK e Institute of Geotechnical Engineering, Nanjing University of Technology, 210009 Nanjing, PR China article info Article history: Available online 25 October 2013 Keywords: Hollow-cylindrical-joint honeycomb Young’s modulus Poisson’s ratio Stress intensity factor Strength abstract In this paper, we constructed a new honeycomb by replacing the three-edge joint of the conventional reg- ular hexagonal honeycomb with a hollow-cylindrical joint, and developed a corresponding theory to study its mechanical properties, i.e., Young’s modulus, Poisson’s ratio, fracture strength and stress inten- sity factor. Interestingly, with respect to the conventional regular hexagonal honeycomb, its Young’s modulus and fracture strength are improved by 76% and 303%, respectively; whereas, for its stress inten- sity factor, two possibilities exist for the maximal improvements which are dependent of its relative den- sity, and the two improvements are 366% for low-density case and 195% for high-density case, respectively. Moreover, a minimal Poisson’s ratio exists. The present structure and theory could be used to design new honeycomb materials. Ó 2013 Elsevier Ltd. All rights reserved. 1. Introduction Honeycomb or foam structures extensively exist in natural materials, e.g. honeycomb [1], cancellous bone in animal skeletons and tree or grass stems [2]. From the mechanical point of view, their mechanical properties (such as light-weight, high-strength and super-tough) are linked to their optimal structures by Nature, which hints people to design different multifunctional materials [3–6]. To this end, a considerable number of scientists and engi- neers invented varieties of porous materials and investigated their mechanical properties. For example, mechanical properties of hier- archical nanohoneycomb or nanofoam materials, for which surface effect was included, were studied [7,8]. It is found that the elastic modulus and strength decrease as hierarchical level number in- creases. Also the sandwich walls with core struts in lattice struc- tures have superior mechanical properties to that with solid walls [9]. Similarly, substituted solid cell walls of the conventional hexagonal honeycomb with equal-mass honeycomb lattice, the Young’s modulus of new structures is optimized and improved by 75% comparing to the conventional hexagonal honeycomb [10]; also, by replacing the three-edge joint of the regular hexago- nal honeycomb with a hollow hexagonal prism, the Young’s modulus of the fractal-like structure is optimized [11]. Besides, replaced the solid cell walls of the conventional hexagonal honeycomb with a equal-mass re-entrant negative Poisson’s ratio honeycomb [12], the Young’s modulus of the new structure is again dramatically improved. Regarding the fracture behavior of honeycomb-like structures, not like the theory for continuum media, the common method is employing the stress field ahead of crack tip, then, performing structural analysis. There exist such models in literatures [13– 15]. Maiti et al. [13] and Choi and Lakes [14] used the conventional singular stress field and the nonsingular stress field for blunt cracks to calculate the axial force acting on the first vertical cell ahead of crack tip, respectively, and both the two models derived fracture strength first and then stress intensity factor. However, Choi and Sankar [15] calculated the axial force and bending mo- ment acting on the first vertical cell ahead of crack tip, and selected an effective portion of crack tip stress field considering the exis- tence of singularity, then, the stress intensity factor was directly obtained. In this paper, by observing natural honeycombs, we found that the cell walls of the natural honeycombs have varying cross- sections, and the thickness reaches a maximum at both ends of cell walls (marked by circle in Fig. 1a). It is speculated that the mo- ments at these ends bear the greatest bending moment, and hon- eybees strengthen their nests by introducing more materials to the weakest points, exactly like the specifications of building de- sign, in which more steel bars and concrete are placed at load- bearing positions in beams. Meanwhile, considering the superior mechanical efficiency of natural porous structures to solid materi- als [8,10,11], we proposed a so-called hollow-cylindrical-joint hon- eycomb (Fig. 1c) by replacing the three-edge joint with a hollow 0263-8223/$ - see front matter Ó 2013 Elsevier Ltd. All rights reserved. http://dx.doi.org/10.1016/j.compstruct.2013.10.025 Corresponding author. Tel./fax: +86 2583792620. E-mail address: [email protected] (Z. Li). Composite Structures 109 (2014) 68–74 Contents lists available at ScienceDirect Composite Structures journal homepage: www.elsevier.com/locate/compstruct

Mechanical properties of a hollow-cylindrical-joint honeycomb · 2021. 1. 20. · nal honeycomb with a hollow hexagonal prism, the Young’s modulus of the fractal-like structure

  • Upload
    others

  • View
    6

  • Download
    0

Embed Size (px)

Citation preview

  • Composite Structures 109 (2014) 68–74

    Contents lists available at ScienceDirect

    Composite Structures

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

    Mechanical properties of a hollow-cylindrical-joint honeycomb

    0263-8223/$ - see front matter � 2013 Elsevier Ltd. All rights reserved.http://dx.doi.org/10.1016/j.compstruct.2013.10.025

    ⇑ Corresponding author. Tel./fax: +86 2583792620.E-mail address: [email protected] (Z. Li).

    Qiang Chen a, Nicola Pugno b,c,d, Kai Zhao e, Zhiyong Li a,⇑a Biomechanics Laboratory, School of Biological Science and Medical Engineering, Southeast University, 210096 Nanjing, PR Chinab Laboratory of Bio-Inspired & Graphene Nanomechanics, Department of Civil, Environmental and Mechanical Engineering, University of Trento, I-38123 Trento, Italyc Center for Materials and Microsystems, Fondazione Bruno Kessler, I-38123 Trento, Italyd School of Engineering and Materials Science, Queen Mary Univerisity of London, Mile End Road, London, E1 4NS, UKe Institute of Geotechnical Engineering, Nanjing University of Technology, 210009 Nanjing, PR China

    a r t i c l e i n f o

    Article history:Available online 25 October 2013

    Keywords:Hollow-cylindrical-joint honeycombYoung’s modulusPoisson’s ratioStress intensity factorStrength

    a b s t r a c t

    In this paper, we constructed a new honeycomb by replacing the three-edge joint of the conventional reg-ular hexagonal honeycomb with a hollow-cylindrical joint, and developed a corresponding theory tostudy its mechanical properties, i.e., Young’s modulus, Poisson’s ratio, fracture strength and stress inten-sity factor. Interestingly, with respect to the conventional regular hexagonal honeycomb, its Young’smodulus and fracture strength are improved by 76% and 303%, respectively; whereas, for its stress inten-sity factor, two possibilities exist for the maximal improvements which are dependent of its relative den-sity, and the two improvements are 366% for low-density case and 195% for high-density case,respectively. Moreover, a minimal Poisson’s ratio exists. The present structure and theory could be usedto design new honeycomb materials.

    � 2013 Elsevier Ltd. All rights reserved.

    1. Introduction

    Honeycomb or foam structures extensively exist in naturalmaterials, e.g. honeycomb [1], cancellous bone in animal skeletonsand tree or grass stems [2]. From the mechanical point of view,their mechanical properties (such as light-weight, high-strengthand super-tough) are linked to their optimal structures by Nature,which hints people to design different multifunctional materials[3–6]. To this end, a considerable number of scientists and engi-neers invented varieties of porous materials and investigated theirmechanical properties. For example, mechanical properties of hier-archical nanohoneycomb or nanofoam materials, for which surfaceeffect was included, were studied [7,8]. It is found that the elasticmodulus and strength decrease as hierarchical level number in-creases. Also the sandwich walls with core struts in lattice struc-tures have superior mechanical properties to that with solidwalls [9]. Similarly, substituted solid cell walls of the conventionalhexagonal honeycomb with equal-mass honeycomb lattice, theYoung’s modulus of new structures is optimized and improvedby 75% comparing to the conventional hexagonal honeycomb[10]; also, by replacing the three-edge joint of the regular hexago-nal honeycomb with a hollow hexagonal prism, the Young’smodulus of the fractal-like structure is optimized [11]. Besides,replaced the solid cell walls of the conventional hexagonalhoneycomb with a equal-mass re-entrant negative Poisson’s ratio

    honeycomb [12], the Young’s modulus of the new structure isagain dramatically improved.

    Regarding the fracture behavior of honeycomb-like structures,not like the theory for continuum media, the common method isemploying the stress field ahead of crack tip, then, performingstructural analysis. There exist such models in literatures [13–15]. Maiti et al. [13] and Choi and Lakes [14] used the conventionalsingular stress field and the nonsingular stress field for bluntcracks to calculate the axial force acting on the first vertical cellahead of crack tip, respectively, and both the two models derivedfracture strength first and then stress intensity factor. However,Choi and Sankar [15] calculated the axial force and bending mo-ment acting on the first vertical cell ahead of crack tip, and selectedan effective portion of crack tip stress field considering the exis-tence of singularity, then, the stress intensity factor was directlyobtained.

    In this paper, by observing natural honeycombs, we found thatthe cell walls of the natural honeycombs have varying cross-sections, and the thickness reaches a maximum at both ends of cellwalls (marked by circle in Fig. 1a). It is speculated that the mo-ments at these ends bear the greatest bending moment, and hon-eybees strengthen their nests by introducing more materials tothe weakest points, exactly like the specifications of building de-sign, in which more steel bars and concrete are placed at load-bearing positions in beams. Meanwhile, considering the superiormechanical efficiency of natural porous structures to solid materi-als [8,10,11], we proposed a so-called hollow-cylindrical-joint hon-eycomb (Fig. 1c) by replacing the three-edge joint with a hollow

    http://crossmark.crossref.org/dialog/?doi=10.1016/j.compstruct.2013.10.025&domain=pdfhttp://dx.doi.org/10.1016/j.compstruct.2013.10.025mailto:[email protected]://dx.doi.org/10.1016/j.compstruct.2013.10.025http://www.sciencedirect.com/science/journal/02638223http://www.elsevier.com/locate/compstruct

  • Fig. 1. (a) Natural honeycomb; (b) conventional regular hexagonal honeycomb; (c) hollow-cylindrical-joint honeycomb.

    Q. Chen et al. / Composite Structures 109 (2014) 68–74 69

    cylinder instead of the hollow hexagonal prism reported in Ref.[11]. The related structure can be regarded as the derivative struc-ture from the family of center-symmetrical honeycombs in Ref.[16]. In particular, the out-of-plane properties of the tetrachiraland hexachiral honeycomb family were well studied by experi-ments and finite element method, and the results showed thatbuckling strength of the honeycombs could be optimized for appli-cations in some fields [16]. For the present structure, its Young’smodulus and Poisson’s ratio was derived basing on Castigliano’stheorem, and its stress intensity factor and strength were calcu-lated by invoking the quantized fracture mechanics (QFM) thanksto the discrete feature of the honeycomb. In the following sections,the mechanical properties of the honeycomb with respect to thoseof the conventional regular hexagonal honeycomb are studied anddiscussed in detail.

    2. Structural theory

    Regarding the relative density of the conventional regular hex-agonal honeycomb (Fig. 1b), it is approximately expressed as

    �qð1Þ ¼ qð1Þqs ¼ 2ffiffi3p tð1Þl� �

    , where, qs is the density of constituent materi-

    als, t(1) is the thickness of cell walls, the superscript (1) denotes theconventional regular hexagonal honeycomb, and its mechanicalproperties dependent of t(1)/l are systematically derived by Gibsonand Ashby [17]. Different from the above expression of relativedensity, here, it is precisely expressed by including a quadraticterm, i.e.,

    �qð1Þ ¼ 13� t

    ð1Þ

    l

    � �þ 2

    ffiffiffi3p� � tð1Þ

    l

    � �:

    As for the relative density of the hollow-cylindrical-joint struc-ture denoted by the superscript (2), we calculate it by a geometri-cal analysis as,

    �qð2Þ ¼ qð2Þ

    qs¼ 2

    3ffiffiffi3p �3 t

    ð2Þ

    l

    � �þ 2ð2p� 3Þ r

    l

    � �þ 3

    � �tð2Þ

    l

    � �ð1Þ

    where t(2) is the thickness of structure’s cell walls, r is the radius ofthe circular joint (Fig. 1c). It is noted that the structure’s geometryrequires 0 < t(2)/r 6 2 and 0 < r/l < 0.5.

    2.1. Young’s modulus

    A representative unit (Fig. 2) is selected and considered as thesum of three components, i.e., semicircle AD, beams BC and DE. Un-der the uniaxially external tensile stress, no rotation and horizontaldisplacement occurs at the end A, thus, the boundary condition ofthe end A is simplified to be guided, see Fig. 2. The force P0 actingon the end C of the beam BC is equivalent to P0 ¼

    ffiffiffi3p

    rð2Þbl=2,where r(2) is the external stress, b is the out-of-plane depth ofthe honeycomb, and l is the length of cell walls.

    For the representative unit ABCDE, according to its geometricalcomponents, the total strain energy stored in it is correspondinglycomposed by three components,

    U ¼ UAD þ UBC þ UDE ð2Þ

    where UAD, UBC and UDE are elastic strain energies stored in thesemicircle AD, beams BC and DE, respectively. Furthermore,the semicircle AD is divided into two subsections: one is AB andthe other BD. Performing force analysis (Fig. 2), we obtain innerforces in the subsection AB: NI = R cos h, VI = R sin h, MI = �MA +Rr(1 � cos h) when 0 < h 6 p/3 and in section BD: NII = �R cosh + P0 sinh, VII = �Rsin h � P0 cos h, MII ¼ MA þM0þP0r

    ffiffi3p

    2 � sin h� �

    � Rrð1� cos hÞ, when p/3 6 h 6 p. Thus, the elasticstrain energy in the semicircle AD is calculated as:

    UAD¼r

    2EsAð2Þ

    Z p3

    0N2I dhþ

    Z pp3

    N2IIdh

    " #þk E

    G

    Z p3

    0V2I dhþ

    Z pp3

    V2IIdh

    " #(

    þAI

    Z p3

    0M2I dhþ

    Z pp3

    M2IIdh

    " #)

    ¼ r2EsAð2Þ

    tð2Þ

    l

    � ��2

    12 p MAl 2þ 4p3 Ml �3P0 rl MAl �2p rl MAl Rh i

    þR 45P0 rl 2�4ð4pþ3 ffiffiffi3p Þ Ml rl þ 34ð1�2kð1þmsÞÞP0 tð2Þl� �2

    � �

    þp2 R2 36 rl

    2þð1þ2kð1þmsÞÞ tð2Þl� �2� �

    þ 8p Ml 2�36P0 Ml rl þP20 4pþ 3 ffiffi3p2� � rl 2h

    þP20 p3þffiffi3p

    8 þ2kð1þmsÞ p3�ffiffi3p

    8

    � �� �tð2Þ

    l

    � �2� �

    8>>>>>>>>>>>>>>>><>>>>>>>>>>>>>>>>:

    9>>>>>>>>>>>>>>>>=>>>>>>>>>>>>>>>>;ð3Þ

    where M ¼ M0 þffiffi3p

    2 P0r is a defined moment.According to Castigliano’s first theorem and boundary condi-

    tions of the end A, the conditions oUAD/oR = 0 and oUAD/oMA = 0hold, then, an equation system with respect to two dimensionlessreaction forces k1 ¼ Rl=M and k2 ¼ MA=M emerges:

    k2 þ C1k1 þ C2 ¼ 0k2 þ C3k1 þ C4 ¼ 0

    �ð4Þ

    where,

    C1¼� 124 rl �1 36 rl 2þð1þ2kð1þmsÞÞ tð2Þl� �2

    � �

    C2¼� 124p rl �1 45 P0 lM� � rl 2�4ð4pþ3 ffiffiffi3p Þ rl þ34ð1�2kð1þmsÞÞ P0 lM� � tð2Þl� �2

    � �C3¼� rlC4¼ 23� 32p

    P0 lM

    � �rl

    h i

    8>>>>>>>>><>>>>>>>>>:

  • Fig. 2. Force analysis in a representative unit. Note that the red curves denote the after-deformed structures. (For interpretation of colour in this figure legend, the reader isreferred to the web version of this article.)

    70 Q. Chen et al. / Composite Structures 109 (2014) 68–74

    solving the system, we obtain k1 ¼ �ðC2 � C4Þ=ðC1 � C3Þ andk2 ¼ ðC2C3 � C1C4Þ=ðC1 � C3Þ. Then, the strain energy by Eq. (3)can be calculated with the known forces P0 and M0. For the present

    structure in Fig. 2, the two forces P0 and M0 satisfy M0 ¼ffiffi3p

    2 P0l2� r

    ,

    i.e., P0 lM ¼ 4ffiffi3p . Then, the strain energy in ABCDE is expressed as:UAD ¼

    P20r

    2EsAð2Þtð2Þ

    l

    � ��2

    3 3p4 k22� 3p2 rl

    k1k2þ p�3

    ffiffiffi3p

    rl

    � �k2

    h i

    þffiffi3p

    4 k1 45rl

    2�ð9þ4 ffiffiffi3p pÞ rl þ 34ð1�2kð1þmsÞÞ tð2Þl� �2� �

    þ3p32 k21 36

    rl

    2þð1þ2kð1þmsÞÞ tð2Þl� �2� �

    þ 3p2 �9ffiffiffi3p

    rl

    þ 4pþ 3

    ffiffi3p

    2

    � �rl

    2þ p3þ ffiffi3p8� ��h

    þ2kð1þmsÞ p3�ffiffi3p

    8

    � ��tð2Þ

    l

    � �2�

    8>>>>>>>>>>>>>>>><>>>>>>>>>>>>>>>>:

    9>>>>>>>>>>>>>>>>=>>>>>>>>>>>>>>>>;ð5Þ

    For the oblique cantilever beam BC, the inner forces are ex-

    pressed as NIII = P0/2, V III ¼ffiffiffi3p

    P0=2;MIII ¼ffiffi3p

    P02

    l2� r

    � x�

    . It isnoted that the moments at the joint B are balanced, i.e.,MIIIð0Þ ¼ MI p3

    þMII p3

    , which proves the correct force analysis

    to some extent. Correspondingly, its elastic strain energy isexpressed as:

    UBC ¼1

    2EsAð2Þ

    Z ð l2�rÞ0

    N2IIIdxþ kEG

    Z ð l2�rÞ0

    V2IIIdxþAI

    Z ð l2�rÞ0

    M2IIIdx

    " #

    ¼ P20r

    2EsAð2Þtð2Þ

    l

    � ��2l

    2r� 1

    � �

    � 3 12� r

    l

    � �2þ 1

    4ð1þ 6kð1þ msÞÞ

    tð2Þ

    l

    � �2" #ð6Þ

    For the vertical cantilever beam DE, only axial deforma-tion occurs, and thus, its elastic strain energy is easily expressedas:

    UDE ¼P20l

    2EsAð2Þð7Þ

    Substituting Eqs. (5)–(7) into Eq. (2), the total elastic strainenergy U is derived. Again, employing Castigliano’s first theorem,the displacement of the beam end C in the force direction is de-rived as,

    DV ¼@U@P0¼ P0l

    EsAð2Þl

    tð2Þ

    � �2f

    rl;tð2Þ

    l

    � �ð8Þ

    where,

    frl;tð2Þ

    l

    � �

    ¼ rl

    � ��

    3 3p4 k22� 3p2 rl

    k1k2þ p�3

    ffiffiffi3p

    rl

    � �k2

    h iþffiffi3p

    4 k1 45rl

    2�ð9þ4 ffiffiffi3p pÞ rl þ 34ð1�2kð1þmsÞÞ tð2Þl� �2� �

    þ3p32k21 36

    rl

    2þð1þ2kð1þmsÞÞ tð2Þl� �2� �

    þ 3p2 �9ffiffiffi3p

    rl

    þ 4pþ 3

    ffiffi3p

    2

    � �rl

    2þ p3þ ffiffi3p8� ��hþ2kð1þmsÞ p3�

    ffiffi3p

    8

    � ��tð2Þ

    l

    � �2�

    8>>>>>>>>>>>>>>><>>>>>>>>>>>>>>>:

    9>>>>>>>>>>>>>>>=>>>>>>>>>>>>>>>;

    þ 12�r

    l

    � �3

    12�r

    l

    � �2þ1

    4ð1þ6kð1þmsÞÞ

    tð2Þ

    l

    � �2" #þ t

    ð2Þ

    l

    � �2

    and thus, the strain e in the representative unit is calculated:

    eV ¼DV

    3l=4¼ r

    Es2ffiffiffi3p

    3l

    tð2Þ

    � �3f

    rl;tð2Þ

    l

    � �ð9Þ

    Finally, the Young’s modulus is obtained:

    Eð2Þ

    Es¼

    ffiffiffi3p

    2tð2Þ

    l

    � �3f�1

    rl;tð2Þ

    l

    � �ð10Þ

    if r/l tends to zero, and the quadratic term (t(2)/l)2 in f(r/l, t(2)/l) is ne-glected due to its smallness (i.e., the shear and axial deformationsare neglected), then, f(r/l, t(2)/l) approach 3/8, and more, Eq. (10) willbe rewritten as E(2)/Es = 2.3(t(2)/l)3, which is the result of the conven-tional regular hexagonal honeycomb reported in Ref. [17].

    2.2. Poisson’s ratio

    In Eq. (3), if we exclude M0 (i.e., let M0 disappear),P0 lM ¼ 2ffiffi3p rl �1

    holds, and the strain energy is now defined as UAD, which isexpressed as:

  • Q. Chen et al. / Composite Structures 109 (2014) 68–74 71

    UAD¼P20r

    2EsAð2Þtð2Þ

    l

    � ��2

    12 rl 2 3p

    4 k22þ p� 3

    ffiffi3p

    2

    � �k2� 3p2 rl

    k1k2

    � �

    þffiffi3p

    2rl

    k1 ð27�8

    ffiffiffi3pÞ rl 2þ 34ð1�2kð1þmsÞÞ tð2Þl� �2

    � �

    þ3p8 rl 2

    k21 36rl

    2þð1þ2kð1þmsÞÞ tð2Þl� �2� �

    þ 10p� 33ffiffi3p

    2

    � �rl

    2þ p3þ ffiffi3p8� �þ2kð1þmsÞ p3� ffiffi3p8� �� � tð2Þl� �2

    26666666666664

    37777777777775

    ð11Þ

    then, UAD;UAD, and M0 should satisfy:

    u ¼ UAD � UADM0

    ð12Þ

    where u is the angular displacement caused by M0, see Fig. 2.According to the structural analysis in Fig. 2, the displacementscaused by the shear and axial forces in part BC are:

    dV ¼2ffiffiffi3p P0r

    EsAð2Þtð2Þ

    l

    � ��2l

    2r� 1

    � �3

    12� r

    l

    � �2þ 3

    2kð1þ msÞ

    tð2Þ

    l

    � �2" #

    dN ¼12

    P0r

    EsAð2Þl

    2r� 1

    � �ð13Þ

    Therefore, the horizontal displacement of the point C is calcu-lated as:

    DH ¼ �ul2� r

    � �sin

    p6� dV sin

    p6þ dN cos

    p6

    ð14Þ

    Finally, the Poisson’s ratio is obtained as:

    m ¼ � eHeV¼ �DH=ð

    ffiffiffi3p

    l=4ÞDV=ð3l=4Þ

    ¼ �ffiffiffi3p

    DHDV

    ð15Þ

    Fig. 3. (a) Preexisting crack in the honeycomb; (b) fracture in vertical cell walls; (c) fract

    2.3. Fracture strength and stress intensity factor

    In this section, we consider two fracture mechanisms in animperfect honeycomb (Fig. 3) as stated in the literature [15]. Oneis that the crack propagates due to the tensile-bending failure ofthe vertical cell wall (Fig. 3b), and the other is due to the bendingfailures of the curved cell wall (B0D0 in Fig. 3c) of the circular jointand the inclined cell walls (B0C0 in Fig. 3d). For the two mecha-nisms, their fracture strengths and stress intensity factors are de-rived, respectively, and the competition between them, like thatin Ref. [18], is discussed.

    2.3.1. Mechanism (1): tensile-bending failure of the vertical cell wallAccording to Choi and Sankar [15], the axial force PE and

    moment ME acting on the first cell wall ahead of crack tip areexpressed as,

    PE ¼Z a

    0

    KIffiffiffiffiffiffiffiffiffi2prp bdr ¼ KIb

    ffiffiffiffiffiffi2ap

    r

    ME ¼Z a

    0

    KIffiffiffiffiffiffiffiffiffi2prp brdr ¼ KIb

    3

    ffiffiffiffiffiffiffiffi2a3

    p

    r ð16Þ

    where a = fl is an effective length, f ¼ 0:411ð�qð2ÞÞ0:308 [15] is adimensionless factor, KI is the mode-I stress intensity factor. If thefailure of the vertical cell wall occurs, the maximum bending stressrmax in the cell wall produced by the forces should equal the tensilestrength rscr of the cell wall materials, i.e.,

    rscr ¼PE

    btð2Þþ 6ME

    bðtð2ÞÞ2¼ Kð2ÞIC;1

    ltð2Þ

    � �2 ffiffiffiffiffi2fpl

    rtð2Þ

    lþ 2f

    � �ð17Þ

    where Kð2ÞIC;1 is the critical stress intensity factor, in which the secondsubscript 1 denotes the first failure mechanism. Then, rearrangingEq. (17), the stress intensity factor is obtained as,

    Kð2ÞIC;1rscr

    ffiffiffiffiffiplp ¼ 1ffiffiffiffiffi

    2fp

    tð2Þl þ 2f

    � � tð2Þl

    � �2ð18Þ

    ure at the point D0 and force analysis; (d) fracture at the point B0 and force analysis.

  • 72 Q. Chen et al. / Composite Structures 109 (2014) 68–74

    The quantized fracture mechanics (QFM) presented by Pugno[19,20], has already been applied to carbon nanotubes and graph-ene. Thanks to the geometric similarity between the present struc-ture and graphene, QFM is employed here to derive thehoneycomb’s fracture strength. If a preexisting crack with length2c exists in the honeycomb (Fig. 3a), the effective length a is de-fined as the fracture quanta, then, the QFM strength of the honey-comb is expressed as rð2Þcr;1 ¼ K

    ð2ÞIC;1=

    ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffipðc þ a=2Þ

    p. Furthermore, we

    assume that the crack length satisfies 2c ¼ nffiffiffi3p

    l, where n is thenumber of cracked cells, then, the strength of the honeycomb isexpressed,

    rð2Þcr;1rscr¼ 1ffiffiffi

    fp

    tð2Þl þ 2f

    � � ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi3p

    nþ fq tð2Þ

    l

    � �2ð19Þ

    2.3.2. Mechanism (2): bending failure of cell wallsBefore discussing the bending failure mechanism, let us study

    the strength of a perfect honeycomb first. Its strength is reachedwhen a cell wall fails due to the maximum bending moment. Wecan see that the moment at the end B (or B

    0due to symmetry) of

    the beam BC (or B0C0) gradually disappears if the ratio r/l ap-proaches 0.5, and the position of the maximum moment switchesfrom the end B (or B

    0) of the oblique cantilever beam BC (or B0C0) to

    the end D (or D0) of the semicircle AD. According to the force anal-ysis with above-obtained reaction forces R and MA, and denotingthe maximum moments at the two positions as MB (or MB0 ) andMD (or MD0 ), respectively, the dimensionless moments areexpressed as:

    MBP0 l¼

    ffiffi3p

    212� rl

    MDP0 l¼

    ffiffi3p

    4 k2 � 2k1 rl

    þ 1

    � 8<: ð20Þ

    Second, for an imperfect honeycomb, the fracture location lo-cates at the points B0 or D0. For these two cases, the force analysisare shown in Fig. 3c and d, respectively. It is noted that 0.5 PE inFig. 3c and d corresponds to P0 in Fig. 2, and thus, substituted P0in Eq. (20) with 0.5PE, the bending moment at the points B0 or D0

    in the cracked honeycomb can be expressed as,

    MB0PEl¼ MBþ0:5MEPEl ¼

    ffiffi3p

    412� rl

    þ f6MD0PEl¼ MDþ0:5MEPEl ¼

    ffiffi3p

    8 k2 � 2k1 rl

    þ 1

    � þ f6

    8<: ð21Þwhere MB and MD, which can be calculated by Eq. (20), are bendingmoments at the points B0 or D0 caused only by the axial force PEthanks to the structural symmetry. For brittle materials, the maxi-mum moment Mmax ¼ 16 rscrbðt

    ð2ÞÞ2, thus, the failure occurs whenMmax ¼maxðMB0 ;MD0 Þ, then, we find,

    Kð2ÞIC;2rscr

    ffiffiffiffiffiplp ¼

    1ffiffiffiffi2fp

    3ffiffi3p

    212�

    rlð Þþf

    � tð2Þl

    � �2MD0 < MB0

    1ffiffiffiffi2fp

    3ffiffi3p

    4 k2�2k1rlð Þþ1ð Þþf

    � tð2Þl

    � �2MD0 > MB0

    8>><>>: ð22Þ

    similarly, the strength is derived by QFM,

    rð2Þcr;2rscr¼

    1ffiffifp

    3ffiffi3p

    212�

    rlð Þþf

    � ffiffiffiffiffiffiffiffiffiffiffiffi3p

    nþfp tð2Þ

    l

    � �2MD0 < MB0

    1ffiffifp

    3ffiffi3p

    4 k2�2k1rlð Þþ1ð Þþf

    � ffiffiffiffiffiffiffiffiffiffiffiffi3p

    nþfp tð2Þ

    l

    � �2MD0 > MB0

    8>><>>: ð23Þ

    considering the limiting case that r/l tends to zero and f = 1, then,

    MD < MB holds and Eq. (22) becomes Kð2ÞIC;2 ¼ 0:307rscr

    ffiffiffiffiffiplp

    tð2Þl

    � �2,

    which is identical to the result of the stress intensity factor of reg-ular hexagonal honeycombs in Ref. [17]; and more, if n = 0, Eq. (23)

    becomesrð2Þ

    cr;2rscr¼ 0:435 tð2Þl

    � �2� 49 t

    ð2Þ

    l

    � �2, which is also the result of the

    compressive strength of perfect regular hexagonal honeycombs inRef. [17].

    Overall, as the ratio r/l varies, the fracture location changes. Forthe whole structure, its strength rð2Þcr can be determined byrð2Þcr ¼minðrð2Þcr;1;r

    ð2Þcr;2Þ according to Eqs. (19) and (23). Correspond-

    ingly, its stress intensity factor Kð2ÞIC is also obtained by minimizingEqs. (18) and (22), i.e., Kð2ÞIC ¼minðK

    ð2ÞIC;1;K

    ð2ÞIC;2Þ.

    3. Results and discussion

    Here, we study the influences of the relative density and param-eter r/l on the Young’s modulus, Poisson’s ratio, fracture strengthand stress intensity factor of the honeycomb normalized by itscounterparts of the conventional regular hexagonal honeycomb.The Poisson’s ratio ms of the constituent material is assumed tobe 0.3. The range for r/l is from 0.1 to 0.45 and �qð2Þ is set in therange from 0.01 to 0.21, then, according to Eq. (1) and r/l, we findt(2)/r varying from 0.001 to 1.741, which satisfies the conditions0 < t(2)/r < 2.

    For the normalized Young’s modulus, we compare the result ofthe present structure by the theory with those of the similar struc-ture studied by experiments and finite element results [11], andparametrically investigate the influences of the relative densityand r/l. The structure in Ref. [11] is controlled by the ratio a/l, inwhich a is the side length of the hexagon (see Fig. 4a). The resultsare reported in Fig. 4. We can see that the present theory (the linein Fig. 4a) agrees well with the experimental (the circle in Fig. 4a)and finite element results (the square in Fig. 4a) even though theyhave different geometries. The former has an optimal value whenr/l � 0.31, which is less than the latter’s optimal value whena/l � 0.33, this is because r is less than a if a hexagon is equivalentto a circle. The parametric study shows that the optimal value, withrespect to the conventional honeycomb, decreases (Fig. 4b) as therelative density increases, and the normalized Young’s modulustends to one, as r/l approaches 0.1 which means that the Young’smodulus of the honeycomb tends to that of the conventional hon-eycomb. Corresponding to the conventional honeycomb, theimprovement of the Young’s modulus is up to 76% when�qð2Þ ¼ 0:01 and r/l = 0.31.

    For the structure’s Poisson’s ratio, the present prediction is alsocompared with the finite element results reported in Ref. [11], anda good agreement is again obtained, see Fig. 5a. If r/l ? 0.1, thePoisson’s ratio tend to the well-known result (i.e., one), and a low-est value m = 0.315 is reached when r/l = 0.38, which is less than thereported value 0.37 when a/l = 0.4 under the common relative den-sity �qð2Þ ¼ 0:06. The reason can be referred to the correspondingdiscussion of Young’s modulus. The parametric study on the Pois-son’s ratio is performed with respect to the relative density and r/l,and the result is plotted in Fig. 5b, which shows that the Poisson’sratio is between 0.313 and 0.996.

    For the fracture strength, the mechanism (1) is absent and themechanism (2) prevails, thus, the fracture location switches fromthe point B0 to the point D0 (Fig. 6a). Interestingly, if the bendingfailure occurs at the point B0, it reaches a minimum whenr/l � 0.25. This is because a smaller r results in a greater cell-wallthickness t(2), which requires a greater bending moment to fail; agreater r results in a smaller moment arm of the beam B0C0, whichalso needs a greater force to fail, and r � 0.25l is in-between.Whereas, if the failure is at the point D0, an increasing r results ina decreasing t(2), thus, the failure bending moment becomes smal-ler and smaller. Moreover, as the number of cracked cells (or cracklength) increases, the normalized fracture strength increases(Fig. 6b). This is because the fracture quanta plays a moreimportant role in the case of a shorter crack which reduces the

  • Fig. 4. (a) Comparison between the present theory of the present structure, experiments and finite element results from the literature [11], when �qð2Þ ¼ �qð1Þ ¼ 0:1; (b)normalized Young’s modulus vs. relative density �qð2Þ and r/l.

    Fig. 5. (a) Comparison between the present theory of the present structure and finite element results from the literature [11], when �qð2Þ ¼ �qð1Þ ¼ 0:06; (b) Poisson’s ratio vs.relative density �qð2Þ and r/l.

    Fig. 6. (a) Normalized fracture strength vs. r/l when n = 7; (b) normalized fracture strength vs. number of cracked unit cells n and r/l when �qð2Þ ¼ �qð1Þ ¼ 0:1.

    Q. Chen et al. / Composite Structures 109 (2014) 68–74 73

    improvement of the fracture strength, and its influence weakens asthe crack length increases. In particular, the maximal improvementof the fracture strength with respect to the conventional honey-comb is up to 300% when r/l = 0.1 and n = 25, and for each n, themaximal improvement is always reached at r/l = 0.1. It is worthmentioning that the critical failure at both points B0 and D0 occurssimultaneously when r/l = 0.33, and its strength is improved by264% when r/l = 0.33 and n = 25.

    Finally, the result of the stress intensity factor is reported inFig. 7. Different from the optimal value of the Young’s modulus

    when r/l = 0.31 (Fig. 4b) and the maximal value of the fracturestrength when r/l = 0.1 (Fig. 6b), the maximal value of the normal-ized stress intensity factor varies from r/l = 0.1 to r/l = 0.33 as therelative density decreases. Addressing this point, we study its crit-ical conditions depicted in Fig. 7a. It shows that the maximal valueappears at both r/l = 0.1 and r/l = 0.33 when �qð2Þ ¼ 0:03; while it isat r/l = 0.33 when �qð2Þ ¼ 0:02 and at r/l = 0.1 when �qð2Þ ¼ 0:04.Therefore, combining Fig. 7b, it can be concluded that the maximalvalue of the normalized stress intensity factor is at r/l = 0.33 if�qð2Þ < 0:03 while at r/l = 0.1 if �qð2Þ > 0:03. Compared to the

  • Fig. 7. (a) Normalized stress intensity factor vs. r/l, note that the three horizontal lines are corresponding to the maximal values of the three cases, respectively; (b)normalized fracture strength vs. relative density �qð2Þ and r/l.

    74 Q. Chen et al. / Composite Structures 109 (2014) 68–74

    conventional regular honeycomb, the stress intensity factor ismaximally improved by 366% when �qð2Þ ¼ 0:01 and r/l = 0.33 andby 195% when �qð2Þ ¼ 0:21 and r/l = 0.1. It is worth mentioning thatif the natural honeycomb has a relative density greater than 0.03,the result illustrates why more silk and wax are centrally locatedat the three-edge joint, according to strength and fracture tough-ness of which the maximal values are at smaller r/l.

    4. Conclusion

    In this paper, we have constructed a hollow-cylindrical-jointhoneycomb, and developed a theory to calculate its Young’s mod-ulus, Poisson’s ratio, fracture strength and stress intensity factor.With respect to the conventional honeycomb, the results showedthat its Young’s modulus can be optimized and comparable to thatin the literature. The smallest Poisson’s ratio is obtained whenr/l � 0.38. For the maximal improvement of its fracture strength,it can be obtained by decreasing the radius of the circular joint.Whereas, a critical relative density 0.03 exists for the maximalimprovement of stress intensity factor, namely, the maximal valueis reached when the ratio r/l equals 0.1 and the honeycomb’s rela-tive density is greater than 0.03; otherwise, the maximal value isobtained when r/l equals 0.33 and the honeycomb’s relative den-sity is less than 0.03. The present structure and theory could beused as a guide to design new honeycomb materials.

    Acknowledgements

    CQ is supported by the Priority Academic Program Develop-ment of Jiangsu Higher Education Institutions (No. 1107037001)and the National Natural Science Foundation of China (NSFC)(No. 31300780). NP thanks the European Research Council (ERCStG Ideas 2011 BIHSNAM, ERC Proof of Concept REPLICA2 2013)and the European Union (Graphene Flagship) for support.

    References

    [1] Zhang K, Duan HL, Karihaloo BL, Wang JX. Hierarchical, multilayered cell wallsreinforced by recycled silk cocoons enhance the structural integrity ofhoneybee combs. Proc Natl Acad Sci USA 2010;107:9502–6.

    [2] Chen Q, Pugno NM. Bio-mimetic mechanisms of natural hierarchical materials:a review. J Mech Behav Biomed Mater 2013;19:3–33.

    [3] Evans AG, Hutchinson JW, Fleck NA, Ashby MF, Wadley HND. The topologicaldesign of multifunctional cellular metals. Prog Mater Sci 2001;46:309–27.

    [4] Zheng Q, Fan H, Liu J, Ma Y, Yang L. Hierarchical lattice composites forelectromagnetic and mechanical energy absorptions. Compos Part B – Eng2013;53:152–8.

    [5] Zheng J, Zhao L, Fan H. Energy absorption mechanisms of hierarchical wovenlattice composites. Compos Part B – Eng 2012;43:1516–22.

    [6] Sun Y, Chen Q, Pugno N. Elastic and transport properties of the tailorablemultifunctional hierarchical honeycombs. Compos Struct 2014;107:698–710.

    [7] Chen Q, Pugno NM. Mechanics of hierarchical 3-D nanofoams. EPL2012;97:26002.

    [8] Chen Q, Pugno NM. In-plane elastic buckling of hierarchical honeycombmaterials. Eur J Mech A/Solids 2012;34:120–9.

    [9] Fan H, Jin F, Fang D. Mechanical properties of hierarchical cellular materials.Part I Analysis. Compos Sci Technol 2008;68:3380–7.

    [10] Taylor CM, Smith CW, Miller W, Evans KE. The effects of hierarchy on the in-plane elastic properties of honeycombs. Int J Solids Struct 2011;48:1330–9.

    [11] Ajdari A, Jahromi BH, Papadopoulos J, Nayeb-Hashemi H, Vaziri A. Hierarchicalhoneycombs with tailorable properties. Int J Solids Struct 2012;49:1413–9.

    [12] Sun Y, Pugno N. In plane stiffness of multifunctional hierarchical honeycombswith negative Poisson’s ratio substructures. Compos Struct 2013;106:681–9.

    [13] Maiti SK, Ashby MF, Gibson LJ. Fracture toughness of brittle cellular solids.Scripta Metall 1984;18:213–7.

    [14] Choi JB, Lakes RS. Fracture toughness of re-entrant foam materials with anegative Poisson’s ratio: experiment and analysis. Int J Fract 1996;80:73–83.

    [15] Choi S, Sankar BV. A micromechanical method to predict the fracturetoughness of cellular materials. Int J Solids Struct 2005;42:1797–817.

    [16] Miller W, Smith CW, Scarpa F, Evans KE. Flatwise buckling optimization ofhexachiral and tetrachiral honeycombs. Compos Sci Technol2010;70:1049–56.

    [17] Gibson LJ, Ashby MF. Cellular solids: structure and properties. 2nded. Cambridge: Cambridge University Press; 1997.

    [18] Chen Q, Pugno NM. Competition between in-plane buckling and bendingcollapses in nanohoneycombs. Europhys Lett 2012;98:16005.

    [19] Pugno NM, Ruoff RS. Quantized fracture mechanics. Philos Mag2004;84:2829–45.

    [20] Pugno NM. Dynamic quantized fracture mechanics. Int J Fract2006;140:159–68.

    http://refhub.elsevier.com/S0263-8223(13)00538-2/h0005http://refhub.elsevier.com/S0263-8223(13)00538-2/h0005http://refhub.elsevier.com/S0263-8223(13)00538-2/h0005http://refhub.elsevier.com/S0263-8223(13)00538-2/h0010http://refhub.elsevier.com/S0263-8223(13)00538-2/h0010http://refhub.elsevier.com/S0263-8223(13)00538-2/h0015http://refhub.elsevier.com/S0263-8223(13)00538-2/h0015http://refhub.elsevier.com/S0263-8223(13)00538-2/h0020http://refhub.elsevier.com/S0263-8223(13)00538-2/h0020http://refhub.elsevier.com/S0263-8223(13)00538-2/h0020http://refhub.elsevier.com/S0263-8223(13)00538-2/h0025http://refhub.elsevier.com/S0263-8223(13)00538-2/h0025http://refhub.elsevier.com/S0263-8223(13)00538-2/h0105http://refhub.elsevier.com/S0263-8223(13)00538-2/h0105http://refhub.elsevier.com/S0263-8223(13)00538-2/h0035http://refhub.elsevier.com/S0263-8223(13)00538-2/h0035http://refhub.elsevier.com/S0263-8223(13)00538-2/h0040http://refhub.elsevier.com/S0263-8223(13)00538-2/h0040http://refhub.elsevier.com/S0263-8223(13)00538-2/h0045http://refhub.elsevier.com/S0263-8223(13)00538-2/h0045http://refhub.elsevier.com/S0263-8223(13)00538-2/h0050http://refhub.elsevier.com/S0263-8223(13)00538-2/h0050http://refhub.elsevier.com/S0263-8223(13)00538-2/h0055http://refhub.elsevier.com/S0263-8223(13)00538-2/h0055http://refhub.elsevier.com/S0263-8223(13)00538-2/h0060http://refhub.elsevier.com/S0263-8223(13)00538-2/h0060http://refhub.elsevier.com/S0263-8223(13)00538-2/h0065http://refhub.elsevier.com/S0263-8223(13)00538-2/h0065http://refhub.elsevier.com/S0263-8223(13)00538-2/h0070http://refhub.elsevier.com/S0263-8223(13)00538-2/h0070http://refhub.elsevier.com/S0263-8223(13)00538-2/h0075http://refhub.elsevier.com/S0263-8223(13)00538-2/h0075http://refhub.elsevier.com/S0263-8223(13)00538-2/h0080http://refhub.elsevier.com/S0263-8223(13)00538-2/h0080http://refhub.elsevier.com/S0263-8223(13)00538-2/h0080http://refhub.elsevier.com/S0263-8223(13)00538-2/h0085http://refhub.elsevier.com/S0263-8223(13)00538-2/h0085http://refhub.elsevier.com/S0263-8223(13)00538-2/h0090http://refhub.elsevier.com/S0263-8223(13)00538-2/h0090http://refhub.elsevier.com/S0263-8223(13)00538-2/h0095http://refhub.elsevier.com/S0263-8223(13)00538-2/h0095http://refhub.elsevier.com/S0263-8223(13)00538-2/h0100http://refhub.elsevier.com/S0263-8223(13)00538-2/h0100

    Mechanical properties of a hollow-cylindrical-joint honeycomb1 Introduction2 Structural theory2.1 Young’s modulus2.2 Poisson’s ratio2.3 Fracture strength and stress intensity factor2.3.1 Mechanism (1): tensile-bending failure of the vertical cell wall2.3.2 Mechanism (2): bending failure of cell walls

    3 Results and discussion4 ConclusionAcknowledgementsReferences