Open access

Universal Quantification of Chemical Bond Strength and Its Application to Low Dimensional Materials

Written By

Bo Xu, Xiaoju Guo and Yongjun Tian

Submitted: 30 November 2010 Published: 01 August 2011

DOI: 10.5772/23792

From the Edited Volume

Graphene Simulation

Edited by Jian Ru Gong

Chapter metrics overview

5,080 Chapter Downloads

View Full Metrics

1. Introduction

The chemical bond strength describes the ability of a chemical bond holding two constituting atoms together. Many physical and mechanical properties of a material, such as melting point, activation energy of phase transition, tensile and shear strength, and hardness, are closely related to the bond strength (Kittle, 2004). However, a universal quantification of bond strength in crystal is hard to be performed due to the lack of effective microscopic parameters to characterize the bond strength. Usually, different characteristic parameters are chosen for different materials. For simple substances and covalent compounds, bond strength is correlated to cohesive energy. For ionic crystals, breaking a chemical bond means overcoming of the electrostatic interaction between anion and cation, which is defined as lattice energy and used to characterize bond strength. Correspondingly, several theoretical definitions of chemical bond strength have been proposed, such as Pauling’s definition for ionic crystals (Pauling, 1929), orbital scaling for covalent crystals (Hultgren, 1932), and two power-law expressions for a variety of materials (Brown & Shannon, 1973; Gibbs et al., 2003). These definitions of bond strength are only valid for some specific types of crystals, and a generalized model of bond strength has not been reached.

In a crystal, bond strength is an intrinsic property of chemical bond, and is regulated by the constitutional atoms as well as the crystal structure. From this viewpoint, bond strength is directly determined by the bond length and the shared bonding electrons. Obviously, greater bond strength would be expected with shorter bond length. The extent of electron sharing, related to the electronegativity difference of bond-forming atoms, is determined by the localized electron density in the binding region. It was found that the greater the localized electron density, the more the effective bonding electrons, and the stronger the bond strength (Gibbs et al., 2003). Most recently, we established a universal semi-empirical quantitative scale to describe the strength of chemical bond in crystals (Guo et al., 2009). The chemical bond strength is defined as the maximum force that a chemical bond can resist under the uniaxial tension along the bond direction which is called tensile unbinding force. We found that the bond strength only relies on two parameters, the bond length and effectively bonded valence electron number of a chemical bond.

In the following, the concept of effectively bonded valence electron number of chemical bond is introduced and the universal quantification model of chemical bond strength is established based on effectively bonded valence electron number and bond length exclusively. The correlation between ideal tensile strength and chemical bond strength is presented. This model allows a convenience determination of chemical bond strength for a variety of materials, ranging from covalent crystals to ionic crystals as well as low dimensional materials. Its application to low dimensional materials, such as graphene, h-BN sheet, and SWNT, are also presented.

Advertisement

2. Methodology

As mentioned in the introduction, the shared bonding electrons in the binding region of two bonded atoms plays a vital role in determining the bond strength. To establish an effective quantification model of bond strength, we must find a practical way to estimate the population of these electrons.

Considering two atoms, A and B, forming a bond in a crystal, the valence electrons are ZA and ZB with coordination numbers of NA and NB, respectively. We first consider a simple case where bonded atom pair possesses totally eight valence electrons. The nominal valence electrons contributed to the A-B bond are nA=ZA/NA and nB=ZB/NBfrom atom A and B, respectively. Larger localized electron density in the binding region would result in stronger bond. However, the bonded electrons localized in the binding region are basically smaller than nA or nB. The Mulliken overlap population of a bond from first-principles calculations can provide a measurement of the bonded electrons (Mulliken, 1955). For example, the calculated population is 0.75 for C-C bonds in diamond and 0.19 for Na-Cl bonds in NaCl crystal. While the determination of population greatly depends on the calculation formalisms (Segall et al., 1996), it is more convenient to find a parameter, which can be easily determined, to serve as an alternative population. Here we propose the effectively bonded valence electron (EBVE) number, nAB, of A-B bond in terms of nA and nB as

nAB=nAnBnA2+nB2E1

The EBVE numbers of diamond (0.707) and NaCl (0.163) are in good agreement with the Mulliken population. Some EBVE numbers of various covalent and ionic crystals are listed in our previous publication (Guo et al., 2009) as well as in the following text.

The ideal A-B bond strength can essentially be defined as an unbinding force of chemical bond, and the physical feature of bond strength for solids becomes more apparent and more accessible than Pauling’s definition or other above-mentioned energy scales. Previously, the power-law behavior of the bond length on bond strength are suggested (Brown & Shannon, 1973; Gibbs et al., 1998, 2003), and the exponential dependence of the resistance of a bond to indenter on the population-related ionicity is emphasized (Gao et al., 2003; He et al., 2005; Li et al., 2008; Simunek & Vackar, 2006). These studies have highlighted the role of the bond length and valance electrons on the strength of a chemical bond. The chemical bond strength, defined as the tensile unbinding force FAB, can then be described quantitatively in terms of the bond length dAB and EBVE number nAB with the formalism,

FAB=CdAB-mexp(knAB)E2

where constants C, m, and k can be deduced from first-principles calculations.

Figure 1.

Typical crystal structures of a) zincblende, b) wurtzite, and c) rock salt with the weakest tensile directions marked.

As soon as the tensile unbinding force of a bond is known, the ideal tensile strength of a crystal is easily accessible. For simple structural crystals shown in Figure 1, generally, the weakest tensile directions, such as 111 of zincblende (ZB), 001of wurtzite (WZ), and 001 of rock salt (RS) structures, are parallel to the axes of the bonds to be broken. Thus, the ideal tensile strength,σhkl , of a crystal along the weakest hkl direction should be correlated with FAB through

σhkl=ShklFAB=CShkldAB-mexp(knAB)E3

where Shkl in the unit of m-2 is the number of the broken bonds per unit area on the (hkl) plane, which has the lowest bond density.

Alternatively but more time-consumingly, the ideal tensile strength of a crystal can be determined from first-principles calculations (Roundy et al., 1999). We acquired the ideal tensile strength of a wide variety of covalent and ionic crystals with a single type of chemical bond with ZB, WZ, or RS structures from first-principles calculations, for which nAB, dAB and Shkl are already known or can be calculated from the experimental values of the lattice parameters (Guo et al., 2009). Three parameters, C, m, and k, in Eqn. 3 can be deduced using the Levenberg-Marquardt method (Levenberg, 1944; Marquardt, 1963) by fitting Eqn. 3 with the crystal tensile strength values from first-principles calculations, leading to,

σhkltheor(Pa)=6.6×10-10ShkldAB-1.32exp(3.7nAB)E4

and

FABtheor(N)=6.6×10-10dAB-1.32exp(3.7nAB)E5

The square of correlation coefficient R2 is 0.996 with the mean absolute fractional deviation of ~7%, indicating that Eqn. 4 and 5 are accurate enough to estimate σhkl and FAB.

Up to now, we are considering two bonding atoms, A and B, with totally 8 valence electrons. The simple formula for tensile unbinding force should be valid for complicate crystals, such as β–Si3N4, α-quartz, and α–Al2O3, where the total valence electrons of two bonding atoms are more than 8. The chemical bonds in these crystals are typical two-electron bond. β–Si3N4 and α-quartz follow the 8-N rule of structure, where N is the valence electron of an atom and 8-N is the corresponding coordination number. Since there are one pair of non-bonding electrons for N in β–Si3N4 and two pairs of non-bonding electrons for O in β-quartz, nN and nO are both equal to 1. In α–Al2O3, The coordination number for Al and O are 6 and 4, respectively, leading to nAl of 0.5 and nO of 1.5. It is then straight forward to calculate the tensile unbinding force (bond strength) with Eqn. 5 for these crystals.

Advertisement

3. Results and discussions

In this section, we will start with the calculation of bond strength in chosen types of materials to understand the relationship of bond strength and crystal structure as well as to trace the relations between the macroscopic properties and bond strength. We will end this section with the bond strength calculations for some low-dimensional materials, such as graphene, h-BN sheet, and SWNT, to demonstrate the effectiveness of our semi-empirical quantification model to these systems.

3.1. IV-A semiconductors

The IV-A semiconductors belong to the family of ANB8-N semiconductors which have similar structures under normal conditions and follow similar phase transition rules under high pressure (Mujica et al., 2003). The procedure presented next can easily be applied to other members of ANB8-N semiconductors.The stable phases of the IV-A materials are graphite, diamond structured Si, Ge, and Sn, and a huge variety of polytypic forms of SiC with comparable energies such as 3C, 2H, 4H, and 6H. The phase transitions of these materials have been itemized previously with fifteen types of crystal structures for IV-A materials (Mujica et al., 2003). ZB, WZ, and RS structures are shown in Figure 1. The other 12 structures are summarized in Figure 2 with the coordination states of the atoms marked. ZB, WZ, RS, graphite, and sh (simple hexagonal) are simple with the marked directions parallel to the axes of broken bonds. Tensile strength can be calculated for materials crystallized in these structures in addition to bond strength.

Several points need to be mentioned for these complicated structures before we discuss the calculation results. Firstly, we give the hexagonal representation instead of the simple rhombohedral representation for r8 structure to show the structure more clearly. Secondly, Si atoms in Imma structure are eight-fold coordinated while Ge atoms in the same structure of are six-fold coordinated (Figure 2f). Thirdly, there are two types of coordination states in the most complicated structure Cmca, although these atoms are identical. We denote the ten-fold coordinated atoms located at 8d sites with white spheres, and the eleven-fold coordinated atoms at 8f sites with gray spheres, as shown in Figure 2g.

The lattice parameters and the calculated bond strengths are listed in Table I. Except the sh-Si and sh-Ge, the arithmetic average of the bond length are given in Table I for structures with different bond lengths. The bond strength of graphite, diamond, lonsdaleite, ZB-Si, WZ-Si, ZB-Ge, ZB-Sn, and ZB-SiC can be referred to our recent publication (Guo et al., 2009).

The bond strength of the IV-A materials as a function of bond length is presented in Figure 3. The tensile unbinding forces are unambiguously grouped by distinct nAB values of 0.943, 0.707, 0.471, 0.354, 0.269, and 0.236, respectively (from top to bottom). As nAB reflects the coordination states, a lower value means a higher coordination number. Bond strength of sequential structures appearing along the pressure induced phase transitions will leap from the higher line to the next higher one. For instance, the experimental phase transition sequence of Si under compression is ZB-Si (nAB=0.707) -Sn (0.471) (Immash) (0.354) Cmca (0.269) (hcpfcc) (0.236). In Figure 3, the bond strength in the same line decreases from left to right in the sequence of C-C, Si-C, Si-Si, Ge-Ge, and Sn-Sn except in the line of nAB=0.269 where the order for Si-Si and Ge-Ge is reversed. The bond strengths of Si-Si and Ge-Ge on other lines are not as distinguishable as those of other bonds. Such an anomaly can be explained by the experimental facts that the radii difference of Si and Ge is not so significant as other elements (rC=0.062 Å, rSi=1.068 Å, rGe=1.090 Å, rSn=1.240 Å). Alternatively, the high transition pressure makes the bond length of Ge comparable to that of Si or even shorter, reminiscent of Cmca (nAB=0.269). Briefly, the bond strength of the chemical bond in this type of materials is determined by the coordination number: the lower the coordination number, the higher the bond strength.

Figure 2.

Crystal structures of IV-A compounds without those in Figure 1. a) Graphite, b) sh, c) bc8, d) r8-h, e) β-Sn, f) Imma, g) Cmca, h) fcc, i) hcp, j) st12, k) bcc, and l) bct.

Table 1.

Lattice parameters and bond strength of IV-A semiconductors.

Figure 4.

Bond strength FABtheor vs. dAB for IV-A compounds.

3.2. ANB8-N ionic crystals

Our simple model for chemical bond strength can easily be applied to ionic crystals, just like the above considered pure covalent and polar covalent ANB8-N materials (Guo et al., 2009). The elemental combinations of IA-VIIA, IB-VIIA, and IIA (except Be)-VIA tend to form ionic crystals. The typical structures of ionic ANB8-N materials are RS and CsCl. ZB and WZ structures are also founded for some IB-VIIA crystals (Shindo et al., 1965). These four structures have been presented in Figure 1 and 2. Other structures, such as R3¯m (H) (Hull & Keen, 1994), P3m1 (Sakuma, 1988), P4/nmm (Liu, 1971), P21/m and Cmcm (Hull & Keen, 1999), can also be found in I-VIIA and IIA-VIA compounds and are shown in Figure 4. The concrete coordination state of each atom in these five structures is clearly shown. The distance between the two neighboring iodine atoms in P3m1-structured CuI is 4.353 Å. Consequently, this compound has a lamellar structure, significantly different from other dense structures.

Figure 5.

Crystal structures of ionic AB compounds. a) P3m1, b) P4/nmm, c) P21/m, d) Cmcm, and e)R3¯m.

In our previous work, the bond strength and uniaxial tensile strength of ten types of rocksalt structured compounds has been calculated (Guo et al., 2009). Crystal parameters and calculated bond strength of other ionic AB compounds are shown in Table 2. For the high pressure phases, the lattice parameters are given under compression. For the monoclinic structures of P21/m-structured AgCl, AgBr and AgI, angles are 98.4, 95.9, and 98.4, respectively. The unbinding tensile force versus bond length for I-VIIA and IIA-VIA compounds listed in Table 2 together with those given in previous work is shown in Figure 5. The bond strength of seventy six chemical bonds locates on seven parallel lines from top to the bottom with decreasing nAB values.

3.3. III-VI crystals

Next step is to treat complicate crystals of AmBn with our semi-empirical quantification model of bond strength. We have discussed the most familiar compounds of β–Si3N4, α-quartz, and α–Al2O3 in Section 2. The material members in this group are tremendous. Here only typical materials with the chemical composition of A2B3 (IIIA-VIA) are chosen for a demonstration. The crystal structures of the compounds as shown in Figure 6 are so complicated that sometimes identical atoms in the same structure have different coordination numbers. The nominal bonding valence electrons of atoms with lower electronegativity are equal to the numbers of outmost electrons, while the determination of the nominal bonding valence electrons of atoms with greater electronegativity is usually challenging.

Figure 6.

Bond strength FABtheor vs. dAB for ionic AB compounds.

For B2O3 of P3121 and B2S3 of lamellar P21/c, all B atoms are three-fold coordinated with O (S) atoms which are bonded to two B atoms. Considering the sp hybridization character of O atoms, the bonding valence electrons of O atoms are set to two. The situation differs from the Cmc2-structured B2O3 where B atoms are four-fold coordinated (Prewitt & Shannon, 1968). In this structure, one third of the O atoms are two-fold coordinated and the other two thirds are three-fold coordinated. For the three-fold coordinated O atoms, the nominal bonding valence electrons number is set to four due to the sp2 hybridized orbital. While for the two-fold coordinated O atoms, only two nominal valence electrons are considered in bonding states and the other four stay at non-bonding states. In the case of Al2S3 with P61 symmetry, one fourth of Al atoms are six-fold coordinated (AlVI) and other three fourths are four-fold coordinated (AlIV), and all S atoms are three-fold coordinated with two AlVI atoms and one AlIV atom (Krebs et al., 1993).

Next structure is Cc for Ga2Se3 where Ga atoms are four-fold coordinated, one-third of Se atoms are two-fold coordinated and the other two thirds are three-fold coordinated with the nominal bonding valence electrons of 2 and 4, respectively (Lübbers & Leute, 1982). For In2O3 with R3¯c symmetry, the coordination number of In and O atoms are six and four, respectively (Prewitt et al., 1969). The C2/m-structured Ga2O3 is even more complicated(Ahman et al., 1996). Half of the Ga atoms are four-fold coordinated (GaIV) while the others six-fold coordinated (GaVI). O atoms in this structure are rather complicated: one third of O atoms are four-fold coordinated to three GaVI atoms and one GaIV atom, another one third are three-fold coordinated to two GaVI atoms and one GaIV atom, and the last one third are three-fold coordinated to one GaVI atoms and two GaIV atom.

The lattice parameters, bond lengths, and bond strength are listed in Table 3. Most of the chemical bonds are two-electron bond. The bond strength versus bond length for these A2B3 and A3B4 compounds are shown in Figure 7 together with the data for Al2O3, -C3N4 and -Si3N4 from the previous work (Guo et al., 2009). Bond strength of the sixteen chemical bonds falls on seven parallel lines. Among all the two-electron bonds, the B-O bonds in P3121-structured B2O3 exhibit the greatest unbinding tensile strength, even higher than that of C-C bond in diamond.

Table 2.

Lattice parameters and bond strength of AB ionic compounds.

Figure 8.

Crystal structures of typical A2B3 and A3B4 compounds. White and gray spheres correspond to elements with higher and lower electronegativity, respectively. a) P3121, b) Cmc21, c) P21/c, d) P61, e) C2/m, f) Cc, g) P3, and h)R3¯c.

An extensive analysis of the ANB8-N materials will give twenty-four nAB values, some of them has been shown in this work. The highest two are 0.943 and 0.856 occurring for three-fold coordinated C-C bond in graphite and three-fold coordinated B-N bond in h-BN, respectively. The common feature of these two bonds is more than two nominal valence electrons distributed on the bond, that is to say the bond order of them is higher than 1. The high bonded valence electron number makes them the strongest bonds to resist tensile force along the direction parallel to the axes of the bond. The next highest nAB of 0.707 is from the four-fold coordinated two-electron bond where each atom contributes one valence electron to the bond, following by 0.643, 0.589, 0.530, 0.474, 0.471 and so on in the sequence. The lowest nAB value is 0.124 corresponds to the Ag-I bond in the Pm3¯m-structured AgI.

Generally, lower coordination number results in higher bonded valence electron number, especially when the valence electron numbers of bonded atoms is the same. However, this argument does not hold when the valence electron numbers are different. For example, the Si atoms forming Si-Si bond in hcp- and fcc-structured Si are twelve-fold coordinated, which is the highest among the chemical bonds discussed above. The bonded valence electron number nAB of these Si-Si bonds are 0.236, higher than the six-fold coordinated I-VIIA bonds (0.165), seven-fold coordinated I-VIIA bonds (0.141), and eight-fold coordinated I-VIIA bonds (0.124). It should be emphasized that the lines denoted with the same nAB in different figures are the same line, for instance, 0.707 in Figure 3 and 7, 0.474 in Figure 5 and 7. As long as the chemical bonds have the same nAB, they will lie on the same line, and bond strength depends strictly on the bonded valence electron number nAB. The relation of coordination number and nAB therefore means that a lower coordination number corresponds to higher bond strength.

Table 3.

Crystal parameters, bond length, and bond strength of A2B3 compounds.

Figure 10.

Bond strength FABtheor vs. dAB for typical A2B3 and A3B4 compounds

3.4. Low dimensional systems

We now apply our semi-empirical model to evaluate the theoretical tensile strength of the low dimensional systems, such as graphene, h-BN sheet, SWNT. As we mentioned before, the highest effective bonded valence electron numbers of 0.943 and 0.856 occur for three-fold coordinated C-C bond in graphite and three-fold coordinated B-N bond in h-BN, respectively. This is consistent with the argument that the sp2 hybridized C-C bond in graphite is the strongest chemical bond (Coulson, 1950). Recently, graphene, a single atomic layer of graphite, has stirred enormous research interests owning to its exceptionally high crystallinity and electronic quality, as well as a fertile ground for applications (Geim & Novoselov, 2007). Experimentally, the mechanical properties of graphene have been identified with atomic force microscope (AFM) nanoindentation, giving a tensile strength of 130 GPa (Lee et al., 2008). Here we predict the theoretical tensile strength σ10theorof graphene in <10> direction using the present model as follows

σ10theor=S10Fcc=Fcc3dcc×δRE6

where R is the thickness of graphene taken as the interlayer separation 3.4 Å of graphite. Fcc and dcc values are listed in Table 4 for graphene and other low dimensional systems. The theoretical tensile strength obtained is 162.7 GPa in the <10> direction, 20% higher than the experimental value.

Whilst graphene has a great application potential in microelectronics, hexagonal boron nitride (h-BN) sheets can find uses as an effective insulator in graphene based electronics. The mechanic properties of h-BN sheets have recently been investigated. With an elastic constant (E2D) of 292 Nm-1, a breaking strain of 0.22, and a thickness of 0.33 nm for a single atomic layer of h-BN (Song et al., 2010), the tensile strength can be deduce to be 97 GPa with the procedure suggested for graphene (Lee et al., 2008). Accordingly, the theoretical tensile strength of h-BN in the <10> direction can be calculated with our semi-empirical model, giving a value of 117 GPa, which is in good agreement with the experimental deduced value.

Carbon nanotube (CNT), with the same covalent sp2 bonds formed between individual carbon atoms as graphen, is one of the strongest and stiffest materials. Direct tensile testing of individual tubes is challenging due to their small size (10 nm or less in diameter). There are several experimental efforts on the mechanical properties of CNT (Yu et al., 2000; Demczyk et al., 2002; Ding et al., 2007; Barber et al., 2005). However, the reported failure stress values display a large variance and are well below the theoretical predicted values in most cases (Ozaki et al., 2000; Mielke et al., 2004), which are attributed to the large number of defects presented on the nanotubes. Accurate measurements of tensile strength require high-quality CNT with well-defined sample parameters, as well as the elimination of measurement uncertainties. Notwithstanding, for zigzag single wall nanotube (SWNT), the theoretical tensile strength along the axial direction can be predicted with our simple model as,

σaxialtheor=SaxialFAB=nFABπ[(Dep+δR2)2-(Dep-δR2)2]E7

where Dep is the diameter of selected nanotube and n is the first index of the chiral vector (n, m) for nanotubes. The theoretical tensile strength of C (10, 0), SiC (10, 0), BN (10, 0), and AlN (10, 0) SWNTs are listed in Table 4.

Table 4.

Parameters, calculated bond strength, and tensile strength for selected low dimensional systems.

Before we end this section, there is one last point need to be mentioned regarding to the dependence of the tensile strength on the direction of the applied tensile stress. A tensile stress tilted away from the axis of a chemical bond would generate a shear component with respect to the bond, and we cannot perform the tensile strength calculation with Eqn. 3 under this circumstance. However, if the shear unbinding strength can be expressed, the ideal strength along any specific direction of a crystal will be accessible. Further studies are therefore highly expected.

Advertisement

4. Conclusion

The bond strength of a variety of chemical bonds are analysized with our semi-empirical unbinding tensile force model. This model proves to be valid for a wide selection of crystals, as well as low-dimensional materials such as graphene and nanotubes. In this model, the chemical bond strength, defined as the tensile unbinding force FAB, can be calculated quantitatively in terms of the bond length dAB and effective bonded valence electron number nAB. It is demonstrated that the bond strength relies strongly on the crystal structure of a solid, in particular, the coordinated states of the bonded atoms. As a result, a chemcial bond formed by identical atom pair would have distinct bond strength in different crystal structures. For example, the the C-C bond strength in graphite is 2.67 times as high as that in diamonds. The definition of unbinding tensile strength provides a more intuitive and general representation of bond strength than those of cohensive energy for covalent crystals and lattice anergy for ionic crystals.

Advertisement

Acknowledgments

This work was supported by NSFC (Grant Nos. 50821001 and 91022029), by NBRPC (Grant No. 2011CB808205).

References

  1. 1. AhmanJ.SvenssonG.AlbertssonJ.1996A reinvestigation of beta-gallium oxide. Acta Crystallographica Section C-Crystal Structure Communications, 526133613380108-2701
  2. 2. BarberA. H.AndrewsR.SchadlerL. S.WagnerH. D.2005On the tensile strength distribution of multiwalled carbon nanotubes. Applied Physics Letters, 8720203106 0003-6951
  3. 3. BrownI. D.ShannonR. D.1973Empirical bond-strength bond-length curves for oxide. Acta Crystallographica Section A, 2952662820108-7673
  4. 4. CoulsonC. A.1950Valence, Oxford University Press, Oxford.
  5. 5. DemczykB. G.WangY. M.CumingsJ.HetmanM.HanW.ZettlA.RitchieR. O.2002Direct mechanical measurement of the tensile strength and elastic modulus of multiwalled carbon nanotubes. Materials Science and Engineering a-Structural Materials Properties Microstructure and Processing, 3341-21731780921-5093
  6. 6. DingW.CalabriL.KohlhaasK. M.ChenX.DikinD. A.RuoffR. S.2007Modulus, fracture strength, and brittle vs. plastic response of the outer shell of arc-grown multi-walled carbon nanotubes. Experimental Mechanics, 47125360014-4851
  7. 7. GaoF. M.HeJ. L.WuE. D.LiuS. M.YuD. L.LiD. C.ZhangS. Y.TianY. J.2003Hardness of covalent crystals. Physical Review Letters, 911015502 0031-9007
  8. 8. GeimA. K.NovoselovK. S.2007The rise of graphene. Nature Materials, 631831911476-1122
  9. 9. GibbsG. V.HillF. C.BoisenM. B.DownsR. T.1998Power law relationships between bond length, bond strength and electron density distributions. Physics and Chemistry of Minerals, 2585855900342-1791
  10. 10. GibbsG. V.RossoK. M.CoxD. F.BoisenM. B.2003A physical basis for Pauling’s definition of bond strength. Physics and Chemistry of Minerals, 3053173200342-1791
  11. 11. GuoX. J.WangL. M.XuB.LiuZ. Y.YuD. L.HeJ. L.WangH. T.TianY. J.2009Unbinding force of chemical bonds and tensile strength in strong crystals. Journal of Physics-Condensed Matter, 2148485405 0953-8984
  12. 12. HeJ. L.WuE. D.WangH. T.LiuR. P.TianY. J.2005Ionicities of boron-boron bonds in B-12 icosahedra. Physical Review Letters, 9410031-9007
  13. 13. HullS.KeenD. A.1994High-pressure polymorphism of the copper(I) halides- a neutron-diffraction study to ~10 GPa. Physical Review B, 509586858850163-1829
  14. 14. HullS.KeenD. A.1999Pressure-induced phase transitions in AgCl, AgBr, and AgI. Physical Review B, 5927507610163-1829
  15. 15. HultgrenR.1932Equivalent chemical bonds formed by s, p, and d eigenfunctions. Physical Review, 406089109070003-1899X
  16. 16. KittleC.2004Introduction to Solid State Physics, Wiley, 978-0-47141-526-8New York.
  17. 17. KrebsB.SchiemannA.LageM.1993Synthesis and crystal-structure of a novel hexagonal modification of Al2S3 with 5-coordinated aluninum. Zeitschrift Fur Anorganische Und Allgemeine Chemie, 61969839880044-2313
  18. 18. LeeC.WeiX. D.KysarJ. W.HoneJ.2008Measurement of the elastic properties and intrinsic strength of monolayer graphene. Science, 32158873853880036-8075
  19. 19. LevenbergK.1944A method for the solution of certain non-linear problems in least squares. Quarterly of Applied Mathematics 21641680003-3569X
  20. 20. LiK. Y.WangX. T.ZhangF. F.XueD. F.2008Electronegativity identification of novel superhard materials. Physical Review Letters, 10023235504 0031-9007
  21. 21. LiuL. G.1971Dense modification of BaO and its crystal structure. Journal of Applied Physics, 4210370237040021-8979
  22. 22. LübbersD.LeuteV.1982The crystal structure of β-Ga2Se3. Journal of Solid State Chemistry, 4333393450022-4596
  23. 23. MarquardtD. W.1963An algorithm for least-square estimation of nonlinear parameters. Journal of the Society for Industrial and Applied Mathematics, 1124314410368-4245
  24. 24. MielkeS. L.TroyaD.ZhangS.LiJ. L.XiaoS. P.CarR.RuoffR. S.SchatzG. C.BelytschkoT.2004The role of vacancy defects and holes in the fracture of carbon nanotubes. Chemical Physics Letters, 3904-64134200009-2614
  25. 25. MujicaA.RubioA.MunozA.NeedsR. J.2003High-pressure phases of group-IV, III-V, and II-VI compounds. Reviews of Modern Physics, 7538639120034-6861
  26. 26. MullikenR. S.1955Electronic population analysis on LCAO-MO molecular wave functions. Journal of Chemical Physics, 2310183318400021-9606
  27. 27. OzakiT.IwasaY.MitaniT.2000Stiffness of single-walled carbon nanotubes under large strain. Physical Review Letters, 848171217150031-9007
  28. 28. PaulingL.1929The principles determining the structure of complex ionic crystals. Journal of the American Chemical Society, 514101010260002-7863
  29. 29. PrewittC. T.ShannonR. D.1968Crystal structure of a high-pressure form of B2O3. Acta Crystallographica Section B-Structural Crystallography and Crystal Chemistry, 2468698740108-7681
  30. 30. PrewittC. T.ShannonR. D.RogersD. B.SleightA. W.1969C rare earth oxide-corundum transition and crystal chemisty of oxides having the corundum structure. Inorganic Chemistry, 89198519930020-1669
  31. 31. RoundyD.KrennC. R.CohenM. L.MorrisJ. W.1999Ideal shear strengths of fcc aluminum and copper. Physical Review Letters, 8213271327160031-9007
  32. 32. SakumaT.1988Crystal-structure of β-CuI. Journal of the Physical Society of Japan, 5725655690031-9015
  33. 33. SegallM. D.ShahR.PickardC. J.PayneM. C.1996Population analysis of plane-wave electronic structure calculations of bulk materials. Physical Review B, 542316317163200163-1829
  34. 34. ShindoK.MoritaA.KamimuraH.1965Spin-orbit coupling in ionic crystal with zincblende and wurtzite structures. Journal of the Physical Society of Japan, 2011205420590031-9015
  35. 35. SimunekA.VackarJ.2006Hardness of covalent and ionic crystals: First-principle calculations. Physical Review Letters, 968085501 0031-9007
  36. 36. SongL.CiL. J.LuH.SorokinP. B.JinC. H.NiJ.KvashninA. G.KvashninD. G.LouJ.YakobsonB. I.AjayanP. M.2010Large scale growth and characterization of atomic hexagonal boron nitride layers. Nano Letters, 108320932151530-6984
  37. 37. YuM. F.LourieO.DyerM. J.MoloniK.KellyT. F.RuoffR. S.2000Strength and breaking mechanism of multiwalled carbon nanotubes under tensile load. Science, 28754536376400036-8075

Written By

Bo Xu, Xiaoju Guo and Yongjun Tian

Submitted: 30 November 2010 Published: 01 August 2011