0% found this document useful (0 votes)
8 views13 pages

Elasticity, Flexibility and Ideal Strength of Borophenes

Uploaded by

rsumit1927
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
8 views13 pages

Elasticity, Flexibility and Ideal Strength of Borophenes

Uploaded by

rsumit1927
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 13

Elasticity, Flexibility and Ideal Strength of Borophenes

Zhuhua Zhang, Yang Yang, Evgeni S. Penev and Boris I. Yakobson*

Department of Materials Science and NanoEngineering, and Department of Chemistry, Rice


University, Houston, Texas 77005, United States
*
email: [email protected]

ABSTRACT: We study the mechanical properties of two-dimensional (2D) boron—borophenes


—by first-principles calculations. The recently synthesized borophene with 1/6 concentration of
hollow hexagons (HH) is shown to have in-plane modulus C up to 210 N/m and bending
stiffness as low as D = 0.39 eV. Thus, its Foppl–von Karman number per unit area, defined as
C/D, reaches 568 nm-2, over twofold higher than graphene's value, establishing the borophene as
one of the most flexible materials. Yet, the borophene has a specific modulus of 346 m2/s2 and
ideal strengths of 16 N/m, rivaling those (453 m2/s2 and 34 N/m) of graphene. In particular, its
structural fluxionality enabled by delocalized multi-center chemical bonding favors structural
phase transitions under tension, which result in exceptionally small breaking strains yet highly
ductile breaking behavior. These mechanical properties can be further tailored by varying the HH
concentration, and the boron sheet without HHs can even be stiffer than graphene against tension.
The record high flexibility combined with excellent elasticity in boron sheets can be utilized for
designing composites and flexible systems.

Keywords: 2D boron, flexibility, strength, phase transition, density functional calculations.

1
Positioned between beryllium and carbon in the periodic table, boron is a key element that
has chemical features of both metal and non-metal. Because of this unique nature, a rich variety
of bonding configurations can form between boron atoms, ranging from normal two-center two-
electron bonds to up to seven-center two-electron bonds,1 enabling boron to form a large number
of allotropes and compounds with most elements. Since the advent of graphene,2 two-
dimensional (2D) materials that are one or several atoms thick reign the current field of materials
research. As boron has demonstrated striking similarity to carbon, forming planar clusters1, 3-8,
cage-like fullerences9-15 and 1D nanotubes 7, 16-21, extensive theoretical efforts have been devoted
to exploring graphene analogues of boron—borophenes22-26. Unlike graphene or hexagonal
boron nitride (h-BN) that have exclusively stable honeycomb lattice, the borophene is predicted
to be polymorphic27 with numerous states near the ground-state energy line, due to a highly
variable network of hollow hexagons (HHs) in a reference triangular lattice. Towards the
synthesis of borophenes, several theoretical works28-29 have suggested using metal substrates,
such as Ag or Cu that may screen only a few of 2D boron structures. Very recently, two
independent experiments have reported the successful synthesis of borophenes on Ag(111) by
using molecular beam epitaxy methods 30-31.

Because of reduced dimensionality, borophenes show qualitatively different properties from


bulk boron. For example, while all existing 3D boron allotropes are semiconducting at standard
conditions, borophenes show metallic characteristic30-32, which also distinguishes them from
other well studied 2D materials usually having bandgaps. Among the properties borophenes may
offer, their mechanical properties are of particular interest and importance. First, since boron is
lighter than most elements, borophenes have lower mass density than other 2D materials. This
opens an intriguing possibility of using borophenes as reinforcing elements for designing
composites, provided that their in-plane stiffness and ideal strength are satisfactorily high.
Second, 2D materials exhibit high levels of flexibility against out-of-plane deformation and are
suitable for fabricating flexible devices. The metallic borophenes are specially potential for
making flexible electrode and contact, both being crucial for nanoelectronics. Fully materializing
the borophenes’ potential into these applications is contingent upon knowing their elastic
properties, ideal strength and structural flexibility. Unlike graphene and h-BN sheet made of
two-electron two-center bonding, borophenes are braided by delocalized multicenter two-
electron bonding33. Little from extensive study of mechanics in graphene and h-BN sheet can be
generalized to borophenes, where qualitatively new phenomena could be anticipated.

Here, we show by first-principles calculations that the experimentally realized borophene,


with a HH concentration of v=1/6 (defined as v=m/N, where m is the number of HHs in a unit
cell of N triangular lattice sites), is a type of material that combines supreme flexibility and
excellent in-plane elasticity. The borophene has an out-of-plane bending stiffness fourfold lower
than graphene but its elastic moduli exceeds one-half of graphene’s value. As a result, borophene
possesses a record high Foppl–von Karman number per unit area34-35 as 568 nm-2, establishing it
as one of the most flexible materials. Despite the extreme flexibility, the ideal strength of
borophene can be over 16 N/m, higher than those of the best known polymer materials, and its
specific modulus is up to 346 m2/s2, rivaling those of graphene. In contrast to other 2D materials,
the borophene can relieve tensile stress by increasing HH concentration, manifested as a series of
tension-induced phase transitions, and thereby withstand a tensile load ~10 N/m at strains as high
as 40%. Moreover, the flexibility, elasticity and strength of borophene can be tailored by varying

2
HH concentration, which may be controlled using different metal substrates in synthesis. These
mechanical properties beyond those of existing 2D materials will promote the applications of
borophenes in nanoelectromechanical systems and flexible electronics.

All the calculations are implemented in Vienna Ab-initio Simulation Package (VASP)
code.36-37 We employ ultrasoft pseudo-potentials for the core region and spin-unpolarized density
functional theory (DFT) based on the generalized gradient approximation of Perdew–Burke–
Ernzerhof 38 functional. The kinetic energy cutoff of the plane-wave expansion is set to be 400
eV. The borophenes are simulated using supercells with a periodic boundary condition. The
vacuum region between two borophene layers in adjacent periodic images is fixed to 15 Å in
order to eliminate spurious interaction. For evaluating the bending stiffness, infinite boron
nanotubes with different diameter are simulated with a primitive tube cell. The Brillouin-zone is
densely sampled for integration, with approximately the same k-point density among different-
sized supercells. Uniaxial tensile strain is applied along x and y directions (see definition later) of
the simulated box. The borophenes are stretched step by step, and, meanwhile, the dimension is
allowed to shrink in the orthogonal direction. To allow possible reconstructions during tension,
we use 1×2 and 3×2 supercells for the v1/6 and triangular sheets, respectively; a part of boron
atoms are displaced along the stretching direction from their ideal position prior to structural
relaxation. Different levels of atomic displacement are considered in order to select the optimal
structure at each applied strain ε. This method has proven reliable for evaluating the strength of
single-walled carbon nanotubes.39 All the atomic positions are fully relaxed using the conjugate-
gradient method until the force on each atom is less than 0.01 eV/Å. The energy barriers for
bond rotation are calculated using the climbing image-nudged elastic-band method.40

The experimentally realized borophenes were characterized to be in different lattice phases.


The triangular (v=0) and v1/6 sheets were proposed as atomic models for a stripped phase30, 41
while the v1/5 sheet was proposed for a phase showing a brick-wall pattern31 in scanning
tunneling microscopy images. The relaxed triangular sheet displays a washboard-like buckling
structure due to an excess of electrons (Figure 1a), whereas the v1/6 and v1/5 sheets are purely
planar because of their high HH concentrations making them both electron-deficient (Figure 1b
and 1c). In addition, we also consider the planar α (i.e. v=1/9) and v1/8 sheets, two of the most
stable structures in vacuum27, and a buckled v1/12 sheet (Figure S1), one of the most stable
structures on Au(111) 29. For ease of discussion, we define the lattice orientation along straight
boron chains as x direction and that perpendicular to the chains as y direction (see Figure 1).

In general, engineering strain can be applied in an arbitrary lattice orientation; to examine


possible anisotropy, we consider two basic cases, that is, the strains applied along the x and y
directions. The calculated stress-strain curves of all borophenes are linearly elastic at small
strains. The elastic response can be either isotropic or anisotropic depending on structural
symmetry. The Young’s modulus, C, and Poisson ratio, σ, of different boron sheets are
summarized in Table I, together with those of other typical 2D materials for comparison. For the
v1/6 sheet, C are 189 N/m and 210 N/m along x and y directions, respectively, lower than 342
N/m for graphene 42; the corresponding σ are 0.15 and 0.17, respectively, almost identical to 0.17
for graphene. Interestingly, C of planar B sheet is quite insensitive to v, varying from 189 N/m to
216 N/m in the x direction and from 208 N/m to 222 N/m in the y direction upon changing v
from 1/5 to 1/9. This insensitivity can be understood by taking borophenes as a binary system
composed of HHs and filled hexagons. With applied in-plane strain, the bond deformation

3
around the HHs is more pronounced than that in the filled hexagons, as shown in Figure 2 for the
v1/6 and α sheets, suggesting that the HHs are mechanically "softer" than the filled hexagons.
Since the applied strain is highly concentrated around the HHs, they dominates the elasticity of
entire sheets and result in the insensitivity of C and σ to v. In the triangular sheet with no HH, the
bond deformation is uniform; C becomes outstandingly high along the ridges, but decreases to
163 N/m across the ridges because the tensile deformation is largely contributed by bond rotation
in this direction.

Since the HHs tend to result in a planar sheet and govern its elasticity, the 2D B sheets with
HHs display insignificant anisotropy. The α sheet is even isotropic due to its high symmetry.
Likewise, the Poisson ratio is almost isotropic upon varying v from 1/5 to 1/9. Interestingly, the
triangular sheet has zero Poisson ratio across the ridges and even negative Poisson ratio when
stretching along the ridges. The negative Poisson ratio results from the Poisson effect along the
out-of-plane direction, that is the applied strain reduces the buckling amplitude of the sheet; in
turn, instead of shrinking, the ridge-to-ridge distance is increased markedly due to a much lower
in-plane modulus across the ridges than along the ridges. The negative Poisson ratio should also
exist in other buckled borophenes with low HH concentrations. Earlier theories have reported
negative Poisson ratio in phosphorene43 and a newly predicted Be5C2 monolayer43-44, both with a
large out-of-plane buckling.

Among the known 2D materials, borophene should be the lightest one since only a few
elements can be simpler than boron. This motivates us to calculate the specific modulus, i.e. the
elastic modulus per mass density C/ρ. Figure 3a summarizes the specific modulus of borophenes.
For the v1/6 sheet, the specific modulus is 346 m2/s2 in the y direction, which reaches 76% of
graphene’s value and 95% of h-BN’s value. Meanwhile, this value is 4-20 times higher than
those of other 2D materials and twofold higher than boron fibers45, opening the possibility of
using borophenes as reinforcing elements. It is of practical importance that the specific modulus
of borophene does not depend much on v, nor on lattice orientation. An exception again arises at
the triangular sheet, which has a record high value of 516 m2/s2 along the ridges but a much
smaller value of 211 m2/s2 across the ridges.

Having revealed the in-plane elasticity of borophenes, we proceed to examine their out-of-
plane bending stiffness D, which is determined by fitting the calculated bending energy per unit
area Eben of a boron nanotube as a function of tube radius r, based on an analytical expression
Eben = Dr -2/2. In this manner, we obtain D = 1.42 eV for graphene, in good agreement with
previous DFT-level values46-50. For the 2D B polymorphs, D turns out to rather small in the x and
y directions, as listed in Table I. In particular, the v1/6 sheet has D as low as 0.39 eV along the
HH rows, smaller than the reported value of any 2D material. The reason for this record small D
is twofold. First, the HHs are aligned into parallel rows, which facilitate bending along the rows
due to decreased bond density therein. Second, the row-to-row spacing in the v1/6 sheet is unique
in a manner that optimizes the length of horizontal B-B bonds in the HH rows and, meanwhile,
minimizes the energy cost of bending more rigid filled hexagons. This point becomes clearer
when comparing structures of the v1/5 and v1/12 sheets, both with HH rows too. In the v1/5 sheet,
the B-B bonds in the HH rows are shorter by 0.03 Å and hence stronger than those in the v1/6
sheet; its D thus rises to 0.56 eV. Yet, in the v1/12 sheet, the HH rows are sparsely spaced and the
wide segments made of filled hexagons are more involved in bending deformation (Figure S1),
which results in an even higher D of 0.92 eV.

4
Generally, decreasing v close to 0 leads to an increase of D and results in more evident
anisotropy, since the corresponding borophenes are increasingly buckled. The triangular sheet
has a maximum buckling of 0.87 Å, which increases D to 1.39 eV along the ridges and up to
4.92 eV across the ridges. This bending stiffness is comparable to other 2D materials with finite
thickness, such as 5.21 eV for phosphorene and 9.14 eV for single-layer MoS2 (close to 9 eV in
an earlier work51).

The combined excellent in-plane stiffness and exceptionally small flexural rigidity render
borophenes as an atomic membrane that is easy to bend yet hard to stretch. The material
parameter characterizing this behavior is the Foppl–von Karman number per unit area, γ, defined
as C/D. Compared to specific modulus, γ varies much more widely from material to material
(Figure 3b). The v1/6 sheet has a highest γ of 568 nm-2, which is over two times higher than 234
nm-2 for graphene and forty times higher than 14 nm-2 for the MoS2 monolayer. The v1/6 sheet
thus represents one of the most flexible materials, yet with considerably high in-plane modulus.
Varying v can significantly modulate γ, which decreases to 377 nm-2 for the v1/5 sheet and 34 nm-
2
for the triangular sheet along the ridges. It is worth mentioning that the v1/5, v 1/8, and α sheets
still have higher γ than graphene, independent of lattice orientation.

The unprecedented flexibility of borophene can be further visualized by wrapping it around


an object and then examining its morphology. For demonstration, we use a BN nanotube, around
which the MoS2, graphene and v1/6 sheet are wrapped, respectively. The 2D materials will bend
either locally or uniformly, depending on their bending stiffness D and curvature r-1. The energy
change per unit area of a 2D material wrapping around a BN nanotube can be expressed as
ΔE=Ebend+Evdw, where Ebend= Dr-2/2 is bending energy and Evdw is van der Waals contribution.
When Evdw> Ebend, the 2D material tends to bend uniformly around the nanotube. Note that Evdw
is a constant but Ebend increases with decreasing tube radius. Therefore, there is a critical tube
radius, below which the 2D material starts to bend locally. By calculating Evdw, the critical tube
radius is determined to be 1.8 nm for MoS2 monolayer and 0.9 nm for graphene but drops to 0.3
nm for the v1/6 sheet. Relaxed structures of MoS2, graphene and v1/6 sheet wrapping around a (7,0)
nanotube (r = 0.3 nm) are shown in Figure S2. In accord with our analysis, both the MoS2 and
graphene bend locally, but the v1/6 sheet is fully wrapped around the nanotube. The demonstrated
extreme flexibility would allow the v1/6 sheet to precisely follow a solid surface with nanoscale
roughness, to an extent that the interspaces between the sheet and surface are minimized to
impede intercalation of external chemical species. This feature may invite potential applications,
such as coating for anticorrosion if a chemically inert capping layer is additionally deposited.

Another important mechanical parameter is the ideal strength of borophenes. Unlike other 2D
materials with covalent bonds, borophenes are made of multi-center bonds, which may bring out
distinctive mechanical response in the elastic limit. The ideal tensile stress versus engineering
strain for the triangular sheet is shown in Figure 4a. At small strains, the sheet exhibits linear
stress-strain relationship, with notable elastic anisotropy. As the strain increases, the stress-strain
behaviors become increasingly nonlinear and show enhanced anisotropy. The peak stress reaches
20.9 N/m along the ridges at ε xx=8.7% and 12.2 N/m at ε yy=14.3% across the ridges. The peak
strengths and critical strains of the triangular sheet are remarkably close to previously reported
values attained by examining the phonon instability52, confirming the validity of our method
based on atomic displacement in predicting the ideal strength. Since other 2D B sheets have
intrinsic negative phonon frequencies that may obscure the appearance of negative frequency

5
induced by strain, we continue to use our method to examine their ideal strength. Figure 4b
presents the results for the v1/6 sheet, whose peak strength is 16.4 N/m at εxx=12.5% across the
HH rows and 15.4 N/m at εyy=10.6% along the HH rows, showing a little anisotropy. As long as
the HHs are included, the ideal strength of borophene changes limitedly upon varying v, as
evidenced by the similar values in the α and v1/8 sheets (Figure 3c). An implication of this result
is that the strength of borophenes can be insensitive to point defects, such as monovacancy and
adatoms (equivalent to changing HHs), in contrast to other 2D materials where these defects
markedly degrade their mechanical performance. While the ideal strengths of borophenes are
inferior to graphene and h-BN sheet53, they are significantly higher than other 2D materials, such
as MoS254 and phosphorene55 (Figure 3c). Of more interest is that all the strengths of borophenes
are reached at exceptionally small critical strains (8%~15%), compared to those of other 2D
materials (Figure 3d). The small critical strains are attributed to an unusual response of the multi-
center bonding at elastic limit, as discussed below.

Materials at critical tensile strain are usually ensued by structural failure. However, this is not
the case for borophenes, which undergo structural phase transitions under tension. Figure 4b
shows that the v1/6 sheet stretched across the HH rows is transformed into a new v1/7 sheet when
εxx > 12.5% and further into an B monolayer comprised of octagons, squares and triangles as ε xx
is increased to 32% (Figure 4b, top inserts). When stretched along the HH rows, the structural
distortion starts to appear in the rows of filled hexagons at εyy=10.6%; then the structure is
transformed to a new B sheet made of triangles and octagons as εyy is increased to 30% (Figure
5b, bottom inserts). The phase transition endows borophenes with high toughness against tensile
loading. For example, the v1/6 sheet can resist a load of 11 N/m and 6 N/m at a tensile strain as
high as 36% applied across and along the HH rows, respectively, at which even the strongest
graphene has been broken in our simulations. The tension-induced phase transition is a common
behavior in the boron sheets with HHs. Our calculations show that the v1/8 and α sheets can all be
transformed into new sheets with lower HH concentrations (Figure S3), giving them a large
loading capability even under extremely high strain (>30%). We have verified that these key
results can be reproduced in simulations using a larger supercell (Figure S4).

The strain-induced phase transition is rare in 2D materials, and it benefits from the variable B
coordination from 3 to 6 as well as structural fluxionality of borophenes enabled by delocalized
multicenter two-electron B-B bonds. The phase transition involves atomic rearrangements via
bond rotations. We thus searched for the transition states of bond rotations in the v1/6 sheet using
nudged elastic-band method. The energy barrier is 2.28 eV for rotating a bond in the HH row
(marked by black thick line) and 1.7 eV for rotating a bond (marked by blue thick line) tilted
with respect to the HH row (Figure 4c). These barriers are over three-fold lower than that (~9 eV)
for Stone-Wales bond rotation in graphene, indicating that the bonds are much easier to flip in
borophenes. Of more significance is that the two barriers can drop to almost zero at ε=10%
applied along and across the HH rows, respectively (Figure 4d). Accordingly, the reaction
energies, i.e. the energy difference between the initial and final states, become highly negative at
ε=10% to help drive the bond flip. We have also noted that in some planar boron clusters the
inner B atoms can rotate against the periphery of the clusters with almost no barrier.56 Another
point worth mentioning is that pentagons and heptagons, which are normal products of a Stone-
Wales bond rotation in carbon materials, relax into triangles or fused hexagons in borophene. In
this manner, the stress created by bond rotation is relieved in the final state, whose energy is
thereby greatly reduced.

6
In contrast, the triangular sheet shows a brittle fracture at a strain of ~14%, above which the
stress sharply drops to zero (Figure 5a). The brittle fracture can be understood from two aspects:
i) Due to excessive electrons, the triangular sheet has out-of-plane buckling, which is decreased
by applied tensile strain; the decreased buckling diminishes the mixing of in-plane and out-of-
plane orbitals and thereby drives more of the excessive electrons to occupy the in-plane
antibonding states, ending up with severely weakened B-B bonds. ii) The dangling bonds along
the cleaved edges are self-passivated due to the variability of B coordination; the self-passivation
is particularly efficient at the cleaved flat B edge (insets in Figure 5a), which naturally exists in
many planar B molecules 1, 4, 17.

In conclusion, we have performed comprehensive first-principles analyses of the mechanical


properties of borophenes, which are shown to combine excellent elasticity, unprecedented
flexibility and high ideal strength. In particular, the borophene with a HH concentration of 1/6
has a bending stiffness fourfold lower than the value of graphene, attributed to optimally spaced
HH rows. Yet, the Young’s modulus of the v1/6 sheet reaches up to 210 N/m, over 60% of the
graphene’s value. Thus, the v1/6 sheet possesses the highest Foppl–von Karman number per unit
area, i.e. the ratio between in-plane modulus and bending stiffness, featuring it as a material that
most easily bends and crumples than it stretches. The high in-plane elasticity of the borophene is
further aided by its low mass density, leading its specific modulus of 346 m2/s2 to be close to that
of graphene. Of more surprise is that this flexible material has an ideal strength of 16 N/m, only
secondary to graphene and h-BN sheet but being several times higher than MoS2 monolayer,
phosphorene and silicene. Being made of delocalized multi-center bondings, the borophene does
not fracture after the peak strength but experiences strain-induced structural phase transitions that
toughen the materials further, to an extent that it still can resist the same levels of loading as
original even at a strain over 35%. All these mechanical properties of the borophene can be
further adjusted by varying the HH concentration. Our findings offer new insight into the
mechanical response of multi-center two-electron bonds and suggest potential applications of
borophenes in making composites and flexible devices.

References
1. Sergeeva, A. P.; Popov, I. A.; Piazza, Z. A.; Li, W.-L.; Romanescu, C.; Wang, L.-S.;
Boldyrev, A. I., Understanding boron through size-selected clusters: structure, chemical bonding,
and fluxionality. Accounts of chemical research 2014, 47 (4), 1349-1358.
2. Novoselov, K. S.; Geim, A. K.; Morozov, S. V.; Jiang, D.; Zhang, Y.; Dubonos, S. V.;
Grigorieva, I. V.; Firsov, A. A., Electric field effect in atomically thin carbon films. Science
2004, 306 (5696), 666-669.
3. Li, W.-L.; Chen, Q.; Tian, W.-J.; Bai, H.; Zhao, Y.-F.; Hu, H.-S.; Li, J.; Zhai, H.-J.; Li,
S.-D.; Wang, L.-S., The B35 cluster with a double-hexagonal vacancy: a new and more flexible
structural motif for borophene. Journal of the American Chemical Society 2014, 136 (35),
12257-12260.
4. Piazza, Z. A.; Hu, H.-S.; Li, W.-L.; Zhao, Y.-F.; Li, J.; Wang, L.-S., Planar hexagonal
B36 as a potential basis for extended single-atom layer boron sheets. Nature communications
2014, 5, 6.
5. Sergeeva, A. P.; Piazza, Z. A.; Romanescu, C.; Li, W.-L.; Boldyrev, A. I.; Wang, L.-S.,
B22– and B23–: All-boron analogues of anthracene and phenanthrene. Journal of the American
Chemical Society 2012, 134 (43), 18065-18073.

7
6. Li, W.-L.; Romanescu, C.; Jian, T.; Wang, L.-S., Elongation of planar boron clusters by
hydrogenation: Boron analogues of polyenes. Journal of the American Chemical Society 2012,
134 (32), 13228-13231.
7. Oger, E.; Crawford, N. R.; Kelting, R.; Weis, P.; Kappes, M. M.; Ahlrichs, R., Boron
cluster cations: transition from planar to cylindrical structures. Angewandte Chemie International
Edition 2007, 46 (44), 8503-8506.
8. Huang, W.; Sergeeva, A. P.; Zhai, H.-J.; Averkiev, B. B.; Wang, L.-S.; Boldyrev, A. I., A
concentric planar doubly π-aromatic B19− cluster. Nature Chemistry 2010, 2 (3), 202-206.
9. Szwacki, N. G.; Sadrzadeh, A.; Yakobson, B. I., B80 fullerene: an ab initio prediction of
geometry, stability, and electronic structure. Physical review letters 2007, 98 (16), 166804.
10. Zhai, H.-J.; Zhao, Y.-F.; Li, W.-L.; Chen, Q.; Bai, H.; Hu, H.-S.; Piazza, Z. A.; Tian, W.-
J.; Lu, H.-G.; Wu, Y.-B., Observation of an all-boron fullerene. Nature Chemistry 2014, 6 (8),
727-731.
11. Sadrzadeh, A.; Pupysheva, O. V.; Singh, A. K.; Yakobson, B. I., The boron buckyball
and its precursors: an electronic structure study. The Journal of Physical Chemistry A 2008, 112
(51), 13679-13683.
12. Li, H.; Shao, N.; Shang, B.; Yuan, L.-F.; Yang, J.; Zeng, X. C., Icosahedral B12-
containing core–shell structures of B80. Chemical Communications 2010, 46 (22), 3878-3880.
13. Wang, L.; Zhao, J.; Li, F.; Chen, Z., Boron fullerenes with 32–56 atoms: Irregular cage
configurations and electronic properties. Chemical Physics Letters 2010, 501 (1), 16-19.
14. Zhao, J.; Wang, L.; Li, F.; Chen, Z., B80 and other medium-sized boron clusters:
Core−shell structures, not hollow cages. The Journal of Physical Chemistry A 2010, 114 (37),
9969-9972.
15. Lv, J.; Wang, Y.; Zhu, L.; Ma, Y., B38: An all-boron fullerene analogue. Nanoscale 2014,
6 (20), 11692-11696.
16. Ciuparu, D.; Klie, R. F.; Zhu, Y.; Pfefferle, L., Synthesis of pure boron single-wall
nanotubes. The Journal of Physical Chemistry B 2004, 108 (13), 3967-3969.
17. Kiran, B.; Bulusu, S.; Zhai, H.-J.; Yoo, S.; Zeng, X. C.; Wang, L.-S., Planar-to-tubular
structural transition in boron clusters: B20 as the embryo of single-walled boron nanotubes.
Proceedings of the National Academy of Sciences of the United States of America 2005, 102 (4),
961-964.
18. Singh, A. K.; Sadrzadeh, A.; Yakobson, B. I., Probing properties of boron α-tubes by ab
initio calculations. Nano letters 2008, 8 (5), 1314-1317.
19. Liu, F.; Shen, C.; Su, Z.; Ding, X.; Deng, S.; Chen, J.; Xu, N.; Gao, H., Metal-like single
crystalline boron nanotubes: synthesis and in situ study on electric transport and field emission
properties. Journal of Materials Chemistry 2010, 20 (11), 2197-2205.
20. Tian, J.; Xu, Z.; Shen, C.; Liu, F.; Xu, N.; Gao, H.-J., One-dimensional boron
nanostructures: Prediction, synthesis, characterizations, and applications. Nanoscale 2010, 2 (8),
1375-1389.
21. Bezugly, V.; Kunstmann, J.; Grundkötter-Stock, B.; Frauenheim, T.; Niehaus, T.;
Cuniberti, G., Highly conductive boron nanotubes: transport properties, work functions, and
structural stabilities. ACS Nano 2011, 5 (6), 4997-5005.
22. Tang, H.; Ismail-Beigi, S., Novel precursors for boron nanotubes: the competition of two-
center and three-center bonding in boron sheets. Physical review letters 2007, 99 (11), 115501.
23. Yang, X.; Ding, Y.; Ni, J., Ab initio prediction of stable boron sheets and boron
nanotubes: structure, stability, and electronic properties. Physical Review B 2008, 77 (4), 041402.

8
24. Wu, X.; Dai, J.; Zhao, Y.; Zhuo, Z.; Yang, J.; Zeng, X. C., Two-dimensional boron
monolayer sheets. ACS Nano 2012, 6 (8), 7443-7453.
25. Zhang, Z.; Penev, E. S.; Yakobson, B. I., Two-dimensional materials: Polyphony in B flat.
Nature Chemistry 2016, 8 (6), 525-527.
26. Lu, H.; Mu, Y.; Bai, H.; Chen, Q.; Li, S.-D., Binary nature of monolayer boron sheets
from ab initio global searches. The Journal of chemical physics 2013, 138 (2), 024701.
27. Penev, E. S.; Bhowmick, S.; Sadrzadeh, A.; Yakobson, B. I., Polymorphism of two-
dimensional boron. Nano letters 2012, 12 (5), 2441-2445.
28. Liu, Y.; Penev, E. S.; Yakobson, B. I., Probing the Synthesis of Two-Dimensional Boron
by First‐Principles Computations. Angewandte Chemie International Edition 2013, 52 (11),
3156-3159.
29. Zhang, Z.; Yang, Y.; Gao, G.; Yakobson, B. I., Two-Dimensional Boron Monolayers
Mediated by Metal Substrates. Angewandte Chemie International Edition 2015, 54 (44), 13022-
13026.
30. Mannix, A. J.; Zhou, X.-F.; Kiraly, B.; Wood, J. D.; Alducin, D.; Myers, B. D.; Liu, X.;
Fisher, B. L.; Santiago, U.; Guest, J. R., Synthesis of borophenes: Anisotropic, two-dimensional
boron polymorphs. Science 2015, 350 (6267), 1513-1516.
31. Feng, B.; Zhang, J.; Zhong, Q.; Li, W.; Li, S.; Li, H.; Cheng, P.; Meng, S.; Chen, L.; Wu,
K., Experimental realization of two-dimensional boron sheets. Nature Chemistry 2016, 8 (6),
563-568.
32. Feng, B.; Zhang, J.; Liu, R.-Y.; Iimori, T.; Lian, C.; Li, H.; Chen, L.; Wu, K.; Meng, S.;
Komori, F., Direct evidence of metallic bands in a monolayer boron sheet. Physical Review B
2016, 94 (4), 041408.
33. Galeev, T. R.; Chen, Q.; Guo, J.-C.; Bai, H.; Miao, C.-Q.; Lu, H.-G.; Sergeeva, A. P.; Li,
S.-D.; Boldyrev, A. I., Deciphering the mystery of hexagon holes in an all-boron graphene α-
sheet. Physical Chemistry Chemical Physics 2011, 13 (24), 11575-11578.
34. Foppl, A., Vorlesungen uber technische Mechanik. B. G. Teubner: 1905.
35. von Karman, T., Festigkeitsproblem im Maschinenbau. Encyklopadie der
Mathematischen Wissenschaften: 1910; Vol. 4.
36. Kresse, G.; Furthmüller, J., Efficiency of ab-initio total energy calculations for metals
and semiconductors using a plane-wave basis set. Computational Materials Science 1996, 6 (1),
15-50.
37. Kresse, G.; Furthmüller, J., Efficient iterative schemes for ab initio total-energy
calculations using a plane-wave basis set. Physical Review B 1996, 54 (16), 11169.
38. Perdew, J. P.; Burke, K.; Ernzerhof, M., Generalized gradient approximation made
simple. Physical review letters 1996, 77 (18), 3865.
39. Dumitrica, T.; Hua, M.; Yakobson, B. I., Symmetry-, time-, and temperature-dependent
strength of carbon nanotubes. Proceedings of the National Academy of Sciences 2006, 103 (16),
6105-6109.
40. Henkelman, G.; Uberuaga, B. P.; Jónsson, H., A climbing image nudged elastic band
method for finding saddle points and minimum energy paths. The Journal of chemical physics
2000, 113 (22), 9901-9904.
41. Dewhurst, R. D.; Claessen, R.; Braunschweig, H., Two‐Dimensional, but not Flat: An
All‐Boron Graphene with a Corrugated Structure. Angewandte Chemie International Edition
2016, 55 (16), 4866-4868.

9
42. Lee, C.; Wei, X.; Kysar, J. W.; Hone, J., Measurement of the elastic properties and
intrinsic strength of monolayer graphene. Science 2008, 321 (5887), 385-388.
43. Wang, Y.; Li, F.; Li, Y.; Chen, Z., Semi-metallic Be 5C2 monolayer global minimum with
quasi-planar pentacoordinate carbons and negative Poisson's ratio. Nature communications 2016,
7.
44. Jiang, J.-W.; Park, H. S., Negative poisson’s ratio in single-layer black phosphorus.
Nature communications 2014, 5.
45. Riley, M.; Whitney, J., Elastic properties of fiber reinforced composite materials. Aiaa
Journal 1966, 4 (9), 1537-1542.
46. Kudin, K. N.; Scuseria, G. E.; Yakobson, B. I., C2F, BN, and C nanoshell elasticity from
ab initio computations. Physical Review B 2001, 64 (23), 235406.
47. Muñoz, E.; Singh, A. K.; Ribas, M. A.; Penev, E. S.; Yakobson, B. I., The ultimate
diamond slab: GraphAne versus graphEne. Diamond and Related Materials 2010, 19 (5), 368-
373.
48. Ma, T.; Li, B.; Chang, T., Chirality-and curvature-dependent bending stiffness of single
layer graphene. Applied Physics Letters 2011, 99 (20), 201901.
49. Koskinen, P.; Kit, O. O., Approximate modeling of spherical membranes. Physical
Review B 2010, 82 (23), 235420.
50. Wei, Y.; Wang, B.; Wu, J.; Yang, R.; Dunn, M. L., Bending rigidity and Gaussian
bending stiffness of single-layered graphene. Nano letters 2012, 13 (1), 26-30.
51. Zou, X.; Liu, Y.; Yakobson, B. I., Predicting dislocations and grain boundaries in two-
dimensional metal-disulfides from the first principles. Nano letters 2012, 13 (1), 253-258.
52. Wang, H.; Li, Q.; Gao, Y.; Miao, F.; Zhou, X.-F.; Wan, X., Strain effects on borophene:
ideal strength, negative Possion’s ratio and phonon instability. New Journal of Physics 2016, 18
(7), 073016.
53. Wu, J.; Wang, B.; Wei, Y.; Yang, R.; Dresselhaus, M., Mechanics and Mechanically
Tunable Band Gap in Single-Layer Hexagonal Boron-Nitride. Materials Research Letters 2013,
1 (4), 200-206.
54. Li, T., Ideal strength and phonon instability in single-layer MoS2. Physical Review B
2012, 85 (23), 235407.
55. Wei, Q.; Peng, X., Superior mechanical flexibility of phosphorene and few-layer black
phosphorus. Applied Physics Letters 2014, 104 (25), 251915.
56. Wang, Y.-J.; You, X.-R.; Chen, Q.; Feng, L.-Y.; Wang, K.; Ou, T.; Zhao, X.-Y.; Zhai,
H.-J.; Li, S.-D., Chemical bonding and dynamic fluxionality of a B15+ cluster: a nanoscale
double-axle tank tread. Physical Chemistry Chemical Physics 2016, 18, 15774-15782.

10
Table I. Calculated Young’s modulus C, Poisson ratio σ, bending stiffness D and mass density ρ
of borophenes and other typical 2D materials. x and y are two principal lattice directions. For a
honeycomb lattice, x and y correspond to armchair and zigzag directions, respectively.

Materials Cx (N/m) Cy (N/m) σx σy Dx (eV) Dy (eV) ρ (kg/m2)


triangular 399 163 -0.23 0 4.76 1.39 7.73×10-7
v1/12 208 161 0.09 0.08 1.33 0.92 6.78×10-7
α 212 212 0.14 0.14 0.79 0.79 6.49×10-7
v1/8 216 222 0.17 0.18 0.74 0.59 6.38×10-7
v1/6 189 210 0.15 0.17 0.56 0.39 6.07×10-7
v1/5 196 208 0.11 0.12 0.52 0.54 5.88×10-7
graphene 342 342 0.17 0.17 1.46 1.46 7.55×10-7
h-BN 273 273 0.21 0.22 0.97 0.97 7.53×10-7
MoS2 125 125 0.25 0.27 9.14 9.14 3.04×10-6
silicene 60 60 0.22 0.22 0.44 0.44 7.24×10-7
phosphorene 90 24 0.16 0.72 5.21 1.35 1.35×10-6

11
a
x

b c

Figure 1. Atomic geometries of borophenes. (a) triangular, (b) v1/6 and (c) v1/5 sheets are shown.

a b 2.4%

1.0%

Figure 2. Maps of bond strains in the (a) v1/6 and (b) α sheets under an applied biaxial strain of
2%.

12
a b 600
x x
y y
400
C /(m /s )

400
2

(nm )
-2
phosphorene
2

phosphorene
graphene

silicene
200

graphene
200

MoS2
h-BN

silicene
MoS2
h-BN
1 1 1 1 1 1 1 1 1 1
0 12 9 8 6 5 0 12 9 8 6 5
0 0
c 40 d
x x
0.3
30 y y
Strength (N/m)

0.2
20

cr
10 0.1

0 0.0
0 1 1 1 0 1 1 1
9 8 6 9 8 6

Figure 3. Elasticity and strength of borophenes. (a) Specific stiffness and (b) Foppl–von Karman
number per unit area (i.e. the ratio between in-plane and bending modulus) of borophenes and
other typical 2D materials. Collected ideal strength (c) and its critical strain (d) in borophenes,
compared with values of other 2D materials. The data for h-BN sheet, MoS2 and phosphorene in
(c) and (d) are obtained from Refs. 53-55, respectively.

a x εxx=8.7% 16% 36% b εxx=10 20% 38%


20 %
20 y
stress (N/m)

stress (N/m)

10
εyy=14% 16%
10 εyy=10 12% 35%
0 %

0 -10
00
0.0 0.1 0.2  0.3 0.0
0 0.1 0.2 0.3  0.4
c 2
d
2.28 2
0
1.7
Er (eV)
E (eV)

-2
1
*

0.65 -4
x
ini: 0eV y
-0.8 -6
0
0.00
0 0.05  0.10

Figure 4. Stress-strain relationship. Tensile stress as a function of uniaxial strain in x and y


directions for the (a) triangular and (b) v1/6 sheets. (c) Minimum energy paths for two possible
Stone-Wales bond rotations in the v1/6 sheet (the rotated bonds are highlighted in thick black and
blue, respectively). The inserts illustrate the atomic structures of initial, transition and final states.
(d) Energy barrier and reaction energy for bond rotation as functions of applied strain. The blue
line in (d) corresponds to the bond rotation shown in blue in (c).

13

You might also like