9
First-principles study of the eects of mechanical strains on the radiation hardness of hexagonal boron nitride monolayers Qing Peng, * Wei Ji and Suvranu De We investigate the strain eect on the radiation hardness of hexagonal boron nitride (h-BN) monolayers using density functional theory calculations. Both compressive and tensile strains are studied in elastic domains along the zigzag, armchair, and biaxial directions. We observe a reduction in radiation hardness to form boron and nitrogen monovacancies under all strains. The origin of this eect is the strain-induced reduction of the energy barrier to displace an atom. An implication of our results is the vulnerability of strained nanomaterials to radiation damage. 1 Introduction Characterization of atomic structures experimentally may be a challenging task for two-dimensional nanomaterials which are only a single atomic layer thick. High-resolution transmission electron microscopes 15 and scanning transmission electron microscopes 6 are usually used to image atomic structures of two-dimensional nanomaterials. To gain a higher resolution, electron beams with shorter wavelengths are needed to illumi- nate a specimen and produce a magnied image. The energetic electrons may knock out the atoms of the material under investigation and cause controlled radiation damage. Radia- tion-induced structural damage thus limits the use of electron microscopes (EM) at high resolutions. Due to monatomic thickness, 2-D nanomaterials are subject to strains under operating conditions. For example, there are strains due to mismatch of lattices or surface corrugation due to the presence of a substrate. 7,8 It has been shown that the strain can substantially aect the band gaps of 2D nanomaterials, and strain can be used to engineer properties of such materials. 911 This is especially of importance to the application of 2D nano- materials for the next generation of electronic devices, such as high frequency eld-eect transistors, 12 graphene-based spintronics, 13 ferromagnetics, 14 antiferromagnetics, 15 nano- waveguides, 16,17 and nanoelectronics. 7 However, strain engi- neering of the controlled radiation damage has not yet been well studied. In this paper, we focus on the study of hexagonal boron nitride (h-BN) a highly promising 2-D nanomaterial. The h-BN monolayer has a honeycomb lattice structure similar to gra- phene, 18 but with a large band gap of 36 eV. 1923 It is chemically inert, with a high thermal conductivity and resistance to oxidation. In recent years, h-BN has been widely applied to micro- and nano-devices, such as insulators with high thermal conductivity, 18 ultraviolet-light emission in optoelectronics, 2428 high-temperature lubricants, 29 and nano-llers in high strength and thermal conductive nanocomposites. 30,31 Recent research has shown that a tunable band gap nanosheet can be con- structed by the fabrication of hybrid grapheneh-BN structures, which opens a new venue for research to the application of h-BN in electronics. 23,32,33 In a previous study of h-BN monolayers, 34 we have noted a nonlinear elastic response prior to mechanical failure around a strain of 0.24. Kim et al. studied controlled radiation damage and edge structures in h-BN membranes and showed that h-BN is more resistant to the electron beam irradiation at 80 kV than graphene. 35 Holmstrom et al. studied carbon nanotubes, graphene, and Si nanowires under irradiation using combined analytical molecular-dynamics and density- functional theory simulations. 36 They reported reduction in radiation hardness with increase in tensile strain explaining this as a result of displacement of atoms to the volume as interstitials. However, this did not clarify how compressive strains would aect radiation hardness and the correspond- ing mechanism. The goal of our present work is to study strain eects on the radiation hardness of h-BN monolayers using ab initio density functional theory calculations. The threshold displacement energy to form h-BN monovacancy congurations is evaluated at both compressive and tensile strain states with three dierent strain modes. The remainder of the paper is organized as follows: the theory and computational method including simulations of controlled radiation damage and ab initio density functional theory are introduced in Section 2, with Results and discussions in Section 3, followed by Conclusion in Section 4. Department of Mechanical, Aerospace and Nuclear Engineering, Rensselaer Polytechnic Institute, Troy, NY 12180, USA. E-mail: [email protected] Cite this: Nanoscale, 2013, 5, 695 Received 20th August 2012 Accepted 12th November 2012 DOI: 10.1039/c2nr32366d www.rsc.org/nanoscale This journal is ª The Royal Society of Chemistry 2013 Nanoscale, 2013, 5, 695703 | 695 Nanoscale PAPER

Nanoscale - Qing Pengqpeng.org/publication/pdf/Peng_2013_nanoscale_radiation.pdf · on the radiation hardness of hexagonal boron nitride ... with a high thermal conductivity and resistance

Embed Size (px)

Citation preview

Nanoscale

PAPER

Department of Mechanical, Aerospace

Polytechnic Institute, Troy, NY 12180, USA.

Cite this: Nanoscale, 2013, 5, 695

Received 20th August 2012Accepted 12th November 2012

DOI: 10.1039/c2nr32366d

www.rsc.org/nanoscale

This journal is ª The Royal Society of

First-principles study of the effects of mechanical strainson the radiation hardness of hexagonal boron nitridemonolayers

Qing Peng,* Wei Ji and Suvranu De

We investigate the strain effect on the radiation hardness of hexagonal boron nitride (h-BN) monolayers

using density functional theory calculations. Both compressive and tensile strains are studied in elastic

domains along the zigzag, armchair, and biaxial directions. We observe a reduction in radiation

hardness to form boron and nitrogen monovacancies under all strains. The origin of this effect is the

strain-induced reduction of the energy barrier to displace an atom. An implication of our results is the

vulnerability of strained nanomaterials to radiation damage.

1 Introduction

Characterization of atomic structures experimentally may be achallenging task for two-dimensional nanomaterials which areonly a single atomic layer thick. High-resolution transmissionelectron microscopes1–5 and scanning transmission electronmicroscopes6 are usually used to image atomic structures oftwo-dimensional nanomaterials. To gain a higher resolution,electron beams with shorter wavelengths are needed to illumi-nate a specimen and produce a magnied image. The energeticelectrons may knock out the atoms of the material underinvestigation and cause “controlled radiation damage”. Radia-tion-induced structural damage thus limits the use of electronmicroscopes (EM) at high resolutions.

Due to monatomic thickness, 2-D nanomaterials are subjectto strains under operating conditions. For example, there arestrains due to mismatch of lattices or surface corrugation due tothe presence of a substrate.7,8 It has been shown that the straincan substantially affect the band gaps of 2D nanomaterials, andstrain can be used to engineer properties of such materials.9–11

This is especially of importance to the application of 2D nano-materials for the next generation of electronic devices, suchas high frequency eld-effect transistors,12 graphene-basedspintronics,13 ferromagnetics,14 antiferromagnetics,15 nano-waveguides,16,17 and nanoelectronics.7 However, strain engi-neering of the controlled radiation damage has not yet beenwell studied.

In this paper, we focus on the study of hexagonal boronnitride (h-BN) – a highly promising 2-D nanomaterial. The h-BNmonolayer has a honeycomb lattice structure similar to gra-phene,18 but with a large band gap of 3–6 eV.19–23 It is chemicallyinert, with a high thermal conductivity and resistance to

and Nuclear Engineering, Rensselaer

E-mail: [email protected]

Chemistry 2013

oxidation. In recent years, h-BN has been widely applied tomicro- and nano-devices, such as insulators with high thermalconductivity,18 ultraviolet-light emission in optoelectronics,24–28

high-temperature lubricants,29 and nano-llers in high strengthand thermal conductive nanocomposites.30,31 Recent researchhas shown that a tunable band gap nanosheet can be con-structed by the fabrication of hybrid graphene–h-BN structures,which opens a new venue for research to the application ofh-BN in electronics.23,32,33

In a previous study of h-BN monolayers,34 we have noted anonlinear elastic response prior to mechanical failure arounda strain of 0.24. Kim et al. studied controlled radiationdamage and edge structures in h-BN membranes and showedthat h-BN is more resistant to the electron beam irradiation at80 kV than graphene.35 Holmstrom et al. studied carbonnanotubes, graphene, and Si nanowires under irradiationusing combined analytical molecular-dynamics and density-functional theory simulations.36 They reported reduction inradiation hardness with increase in tensile strain explainingthis as a result of displacement of atoms to the volume asinterstitials. However, this did not clarify how compressivestrains would affect radiation hardness and the correspond-ing mechanism.

The goal of our present work is to study strain effects on theradiation hardness of h-BN monolayers using ab initio densityfunctional theory calculations. The threshold displacementenergy to form h-BN monovacancy congurations is evaluatedat both compressive and tensile strain states with threedifferent strain modes. The remainder of the paper is organizedas follows: the theory and computational method includingsimulations of controlled radiation damage and ab initiodensity functional theory are introduced in Section 2, withResults and discussions in Section 3, followed by Conclusion inSection 4.

Nanoscale, 2013, 5, 695–703 | 695

Nanoscale Paper

2 Computational method2.1 Controlled radiation damage

Atoms in the h-BN monolayer may be knocked out by electronbeams and leave vacancies at the boron or nitrogen sites. Theminimum amount of energy that is transferred from the kineticenergy of the electrons to knock out an atom is referred to asthreshold displacement energy Ed (TDE), a characteristic of theradiation hardness of a material. The TDE can be measuredexperimentally by utilizing electron irradiation sources toproduce isolated point defects (vacancies and interstitials).Optical spectroscopy and transmission electron microscopy aretwo techniques used tomeasure the TDE.37 Optical spectroscopycan uniquely monitor the behavior of a particular defect (e.g.,anion vacancy for example). The drawback of this technique,however, is that the signal may not be visible unless displace-ment damage occurs.38 In situmeasurement techniques, such astransmission electron microscopy, are used to monitor defectcluster formation or amorphization. Since it is generallyassumed that visible damage requires displacement of defectsfrom all of the sub-lattices, the TDE of only the most massiveatomic species may be measured accurately using such tech-niques.39 In contrast to experimental techniques, the advantageof simulation is the ability to compute the TDE for all atoms. Inour present work, we employ simulation techniques to deter-mine the TDE to form monovacancies in h-BN monolayers.

The likelihood of radiation damage caused by an incomingelectron beam is quantied by a radiation damage cross-section, a quantity dependent on the energy of incoming elec-tron beams and the TDE of studied materials. The radiationdamage cross-section for isotropic knock-on radiation damagefor a deformed system with a Lagrangian strain of h can becalculated as:40

sh

d ¼Z2e4

�1� b2

�16p302m0

2c4b4

�t� 1� b2ln ðtÞ þ pZab

�2t1=2 �lnðtÞ � 2

��;

(1)

where Z is the atomic number of the displaced atom, e is theelectron charge, m0 is the electron mass, 30 is the dielectricpermittivity of free space, c is the speed of light, a is the nestructure constant, and b ¼ v/c with v the speed of electrons inelectron beams, whose energy is E. According to the theory ofrelativity, b is calculated as

b ¼ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi1�

m0c

2

E þm0c2

2s

: (2)

In eqn (1), t ¼ Tm/Ehd, where Tm is the maximum transferred

energy in a scattering event given by:

Tm ¼ E þ 2mc2

1þ ðM þm0Þ2c22ME

; (3)

where M is the atomic mass of the displaced atom. Ehd is thedisplacement threshold energy2,41 for a deformed system with astrain of h, which is a minimum amount of kinetic energyneeded to displace the atom away from the original sitepermanently.

696 | Nanoscale, 2013, 5, 695–703

Molecular dynamics simulations are normally used to esti-mate Ehd by simulating radiation-induced displacement cascadeprocesses.2,41 However, such computations may overestimatethe threshold displacement energies.35 In our research, therelative change of the radiation resistance to different strainsfor h-BN materials is the major concern. For this, a staticapproach35 can be used to estimate Ehd, which can be regarded asthe lower bound of the threshold displacement energies:

Ehd ¼ Eh

v + Ea � EhBN, (4)

where Ehv is the total energy of the system under strain hwith thevacancy v, Ea is the energy of the displaced single atom in a cubewith a side length of 15 A, and EhBN is the total energy of thesystem under strain h without any vacancy.

2.2 Density functional theory calculations

We used a system in which the supercell contains 112 latticesites in one plane to model the h-BN monolayer with dimen-sions 17.601 A � 17.421 A. The x axis (direction 1) is along thezigzag direction and the y axis (direction 2) is along thearmchair direction. In such a conguration, the deformationsalong the two directions are similar. Periodic boundary condi-tions were applied along the x and y directions. The dimensionsof the supercell were selected to ensure that the interactionsbetween the vacancies and their self-images were negligible.

Due to anisotropy of their atomic structure, h-BNmonolayersexhibit different mechanical properties along different spatialdirections. Therefore, we applied uniaxial strains along thezigzag (z), armchair (a), and biaxial (b) directions in the plane ofthe nanostructure. Both compression and tension wereconsidered. We studied the system response under Lagrangianstrains ranging from �0.1 to 0.1, with an increment of 0.02 foreach step. This range was selected so that deformations werewithin the elastic domain. The strains were applied by changingthe domain size while keeping the number of atomsunchanged. The initial positions of all atoms at the strainedstates were homogeneously rescaled according to the appliedstrain. The structural optimization was achieved by minimiza-tion of the total energy using the conjugated gradient method.

DFT calculations were carried out with the Vienna Ab initioSimulation Package (VASP),42 which is based on the Kohn–ShamDensity Functional Theory (KS-DFT)43 with the generalizedgradient approximations as parameterized by Perdew, Burke andErnzerhof (PBE) for exchange–correlation functions.44 The elec-trons explicitly included in the calculations are the (2s22p1) elec-trons of boron and (2s22p3) electrons of nitrogen. The coreelectrons (1s2) of boron and nitrogen were replaced by theprojector augmented wave (PAW) and pseudo-potentialapproach.45 A plane-wave cutoff of 400 eV was used in all the cal-culations.The calculations were performed at zero temperature.

The atomic structures of all the deformed and undeformedcongurations were obtained by fully relaxing a 112-atom-unitcell. The criterion to stop the relaxation of the electronic degreesof freedom was set by the total energy change to be smaller than0.00001 eV. The optimized atomic geometry was achievedthrough minimizing the Hellmann–Feynman forces acting on

This journal is ª The Royal Society of Chemistry 2013

Paper Nanoscale

each atom until the maximum forces on the ions were smallerthan 0.01 eV A�1.

The irreducible Brillouin Zone was sampled with a gamma-centered 5 � 5 � 1 k-mesh, ensuring the energy convergence to1meV, reducing the numerical errors caused by the strain of thesystems. The initial charge densities were taken as a superpo-sition of atomic charge densities. There was a 15 A thickvacuum region to reduce the inter-layer interaction to model thesingle layer system. To eliminate the articial effect of the out-of-plane thickness of the simulation box on the stress, we usedthe second Piola–Kirchhoff stress34 to express the 2D forces perlength with units of N m�1.

Fig. 1 Stress–strain responses, geometries, and strain energies for uniaxialstrains in the armchair a and zigzag z directions, and biaxial b strains for (top)pristine h-BN, (middle) B-vacancy (VB), and (bottom) N-vacancy (VN) systems.

3 Results and discussion3.1 Elastic response

The lattice constant of the equilibrated h-BN monolayer in thestrain free state was found to be 2.512 A, which was in goodagreement with the experiment (2.51 A).46 When strains wereapplied, all the atoms were fully relaxed within the plane.

The second P–K stress versus Lagrangian strain relationshipsof the three strainmodes are shown in the top panel of Fig. 1. Formode z, the stress in the zigzag direction (Sz

1) increases from�42.8 to 20.4 GPa. Due to xed boundary conditions, the stress inthe armchair direction (Sz

2) is not zero, but is small compared toSz1, increasing from �7.2 to 6.3 GPa. The stress–strain relation-

ships in mode a are very similar to those in mode z, indicatingisotropic mechanical properties in h-BNmonolayers. It should benoted that although mass density differs along the lines in thezigzag and armchair directions, Sz

1 and Sa2 show similar

responses to applied strains. This could be explained by theenergy changes that are averaged by the cross-sectional areaperpendicular to the uniaxial strain direction. For the samereason, Sz

2 and Sa1 act similarly to applied strains. In the case of

tension where strain h > 0, Sz1, S

a2, and Sb

1 all respond similarly,while it is not true for h < 0. The direction dependent stress–strainresponses are a result of anisotropy of the h-BN atomic structure.The asymmetrical stress–strain curves around zero strain reectthe anharmonic responses to compression and tension.

It was found that the system's energy increases when strainis applied. Here we dened the total strain energy as Es ¼ (Etot �E0), where Etot is the total energy of the strained system, and E0is the total energy of the reference state which is strain-free. Esresponds similarly to the uniaxial strain in the z and a direc-tions, but the increase is greater in biaxial straining (Fig. 1). Forsmall strains, a factor of two is held as expected. Es is asym-metrical for compression (h < 0) and tension (h > 0) for all threecases. Es deviates from the quadratic relationship with h beyonda strain of 0.04 in the three tested deformations.

The two-dimensional Young's modulus is dened as

Y2D ¼ 1A0

v2Es

vh2

����h¼0

, where A0 is the equilibrium surface area at a

strain-free state. This quantity is computed as Y2Dh-BN¼ 301 Nm�1.

If we take the interlayer separation as 3.3 A,47 the correspondingthree-dimensional Young's modulus is Y3Dh-BN ¼ 0.99 TPa.This result is in agreement with the experiment Y2Dh-BN,exp z 220–

This journal is ª The Royal Society of Chemistry 2013

510 N m�1, which is obtained for few-layered h-BN.48 Comparedto graphene, (Y2DG ¼ 340 � 50 N m�1 experimentally49), the stiff-ness of h-BN is smaller (89%) but still remarkably high.

A similar study was performed for B-vacancy (VB) congu-ration (Fig. 1, middle). The stress in the zigzag direction (Sz

1)increases from �38.8 to 19.2 GPa, and Sz

2 from �7.4 to 6.2 GPa.

Nanoscale, 2013, 5, 695–703 | 697

Nanoscale Paper

Sb1 varies from �56.7 to 21.3 GPa within the strain range, which

slightly differs from Sb2 which varies from �58.7 to 21.3 GPa.

The Es values of VB are quite similar to those of ideal h-BN. Thetwo-dimensional Young's modulus is Y2DVB ¼ 273 N m�1, smallerthan that of the pristine h-BN. The inset is the geometry of asingle hexagonal ring in the strain-free state. The zoomed-in

Fig. 2 Contour plot of the von Mises atomic strains of the configurations under tharmchair, (c) zigzag, and (e) biaxial. The VN counterparts are (b), (d) and (f). The bla

698 | Nanoscale, 2013, 5, 695–703

geometry plots around the vacancy illustrate the strain status inthese three deformation modes. For uniaxial strain along thezigzag direction, under a compressive strain of hz ¼ �0.1, thetwo N atoms are pushed together to form a N–N bond, with abond length of 1.444 A. For a tensile strain of hz ¼ 0.1, the two Natoms in the previous N–N bond are 3.46 A apart.

e 0.1 global strains. The VB counterparts under three deformation modes are (a)ck circles denote boron atoms and the red hexagons denote nitrogen atoms.

This journal is ª The Royal Society of Chemistry 2013

Paper Nanoscale

Similar stress–strain relationships and elastic energy curvesare observed in the N-vacancy conguration (Fig. 1, bottom).For example, Sz

1 increases from �41.8 to 19.2 GPa within thestrain range and Sb

1 ¼ Sb2 varies from �59.6 to 21.4 GPa. The

two-dimensional Young's modulus calculated from the strainenergy curve is Y2DVN ¼ 285 N m�1, smaller than that of the ideal

Fig. 3 Distribution of the difference in electronic charge density rdiff (in units of e Afor VB. The VN counterparts are (b), (d) and (f). The black circles denote boron atom

This journal is ª The Royal Society of Chemistry 2013

h-BN but bigger than that of VB. For uniaxial strain along thezigzag direction, under a compressive strain of hz ¼ �0.1, thetwo B atoms are pushed together to form a B–B bond with abond length of 1.73 A. For a tensile strain of hz ¼ 0.1, the two Batoms in the previous B–B bond are 3.17 A apart. When strainswere applied along the armchair direction, at a compressive

�3) under deformation along the armchair (a), zigzag (c) and biaxial (e) directionss and the red hexagons denote nitrogen atoms.

Nanoscale, 2013, 5, 695–703 | 699

Nanoscale Paper

strain of ha ¼ �0.1, one boron atom displaces towards thevacancy site, bonded with two B atoms, forming two hexagonalrings and two pentagonal rings. The central B atom has two B–Nbonds and two N–N bonds, with bond lengths of 1.45 and1.82 A, respectively. When biaxial strains were applied, at acompressive strain of hb ¼�0.1, the three B atoms were bondedtogether, forming three pentagons around the vacant site. TheB–B bond lengths are 1.69 A. The bond lengths of these three Natoms with nearby B atoms are 1.35 A. At a tensile strain of hb ¼0.1, the B–B bonds are broken. The bond lengths of these threeB atoms with nearest N atoms are 1.52 A.

In this study, the presence of monovacancy (0.9% of theoverall lattice sites) reduces the Young's modulus by 9% and 5%when it corresponds to a B or N atomic site, respectively.

Fig. 4 Displacement threshold energy Ed,v varies with strain h in the threedeformation modes, in the plane of h-BN monolayer forming B or Nmonovacancy.

3.2 Distribution of strain and electronic charges

The spatial distribution of strains in a defect system is ofimportance in understanding the formation and evolution ofdefects. Due to the presence of monovacancy, applied strainswere not uniformly distributed in h-BN monovacancy congura-tions. To study strain eld near defects, atomistic strain tensors50

were obtained in the deformed systems with monovacancy. As atypical representative of the atomistic strain tensor, the vonMises equivalent strain 3eq., a second rank invariant51 of thestrain tensor 3, is of special interest and is dened as

3eq: ¼ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi3

2Trð3� 3+IÞ2

r; 3+ ¼ 1

3Tr3 (5)

where I is the unity tensor and 3 is a rst rank invariant of thestrain tensor 3.

Compared to local atomistic strains, the applied strain h isglobal in nature. The distribution of local von Mises equivalentatomistic strains in VB and VN systems at a global compressivestrain of 0.1 is plotted in Fig. 2. Large von Mises equivalentatomistic strains were observed around vacancies. The straineld is anisotropic, and is dependent on the amount and thedirection of the applied strain, as well as the vacancy species.

The presence of vacancies also affects the distribution ofelectron charge densities. We calculated the difference in elec-tron charge densities between the h-BN monovacancy systemand the ideal system, denoted as rdiff ¼ rvac � rideal. The resultsare plotted in Fig. 3.

As expected, the largest rdiff appeared near defects. Thedistribution of rdiff differed at different strains with differentmonovacancy species. This observation is consistent with thedistribution of von Mises equivalent atomistic strains as shownin Fig. 2.

3.3 Radiation hardness

Radiation hardness is measured by the displacement thresholdenergy Ed,v, which was calculated based on eqn (4) for h-BNmonolayers under strains. Strains were applied along thezigzag, armchair, and biaxial directions. Values of Ed,v areobtained and plotted as a function of strain h in the threedeformation modes as shown in Fig. 4.

700 | Nanoscale, 2013, 5, 695–703

Without any strain, a higher threshold displacement energywas obtained for the formation of boron monovacancy than theformation of nitrogen monovacancy in the h-BN monolayer(17.08 eV for B and 16.04 eV for N), which is in agreement with aprevious study.35 As both compressive and tensile strains wereapplied, there was a general reduction in radiation hardness ofh-BN monolayers to form both B-vacancy and N-vacancycongurations. The reduction rate differed at strains alongdifferent directions.

The calculation of threshold displacement energy was per-formed assuming that all the structures were fully relaxed. Suchrelaxed structures in the static method provide a good approxi-mation in determining the threshold displacement energy basedon dynamics methods or experimental measurements. Thefollowing two reasons can justify this approximation: (1) theforces between the displaced atom and its adjacent atoms wereopposite and instantaneous. A displaced atom can travel a fewangstroms away from its original equilibrium site prior tocollision. Adjacent atoms, however, can move at a distance of anorder of magnitude shorter to achieve relaxation. The relaxationprocedure can be regarded as instantaneous compared to themovement of the displaced atom. (2) Compared to the manypossible unrelaxed congurations, the relaxed conguration wasunique and determined. The energy difference between the tworelaxed structures, before and aer displacement of the atom,provided a lower limit of the threshold displacement energy.

Our result of the reduction in radiation hardness as tensilestrains were applied agrees well with previous studies byHolmstrom et al., where only response to tensile strains wasexamined.36 The authors attributed such behavior to twoeffects: the increase of space between atoms due to strainingallowing the displaced atoms to settle as interstitials andweakening of bonds as a result of stretching. Such explanation,however, cannot be used for reduction in hardness atcompressive strains.

An energy perspective can better explain the general reduc-tion in radiation hardness, as observed in the present work.

This journal is ª The Royal Society of Chemistry 2013

Paper Nanoscale

Increased strain energy, irrespective of tension or compression,lowers the energy barrier of displacing an atom. Hence, lessenergy is required from the radiation. Therefore, when thesystem is subjected to strains, radiation hardness is reduced.

The threshold displacement energy showed strong nonline-arity with change in strain. The threshold displacement energyis smaller in the compressive than in the tensile strain regimes

Fig. 5 Contour plot of the cross-section shd (in units of barn) with respect to the elearmchair, (b) zigzag, and (c) biaxial. The VN counterparts are (d), (e) and (f).

This journal is ª The Royal Society of Chemistry 2013

for displacements of both boron and nitrogen atoms. Theenergy interpretation discussed above can well explain thedifference of threshold displacement energy betweencompressive and tensile strains. When a compressive strain isapplied, the distances of atoms are shortened and the repulsiveforces rise. For the same amount of strain, the repulsive forcesincrease much faster in compressive strain than the attractive

ctron beam energy and the strain h. The three deformation modes for VB are (a)

Nanoscale, 2013, 5, 695–703 | 701

Nanoscale Paper

forces in tensile strain. Thus systems in the compressive statecan store higher strain energy than in the tensile state, espe-cially for biaxially strained systems, with the same amount ofstrain. We can conclude that the system with higher strainenergy is more vulnerable to radiation damage. As strains areapplied, the input energy is utilized in bond formation,strengthening, attenuation and rupture. The stored energiesdiffer at various strain modes and congurations in the VB andVN systems. This may be the underlying reason that radiationhardness is different for forming a B monovacancy and a Nmonovacancy in different strain modes.

We also examined the threshold displacement energiesusing unrelaxed vacancy congurations (unrelaxed-TDEs). Theunrelaxed-TDE is the sum of TDE and the vacancy relaxationenergy, which is an energy gain during the relaxation aer theremoval of the atom. The vacancy relaxation energy is also afunction of applied strains. As a result, unrelaxed-TDEs are amore complex function of the applied mechanical strains thanTDEs. Under uniaxial compression along either the zigzag orarmchair direction, the unrelaxed-TDEs increase with theamount of the compressive strains, opposite to the decrease ofTDEs under the same circumstances, for both B and N mono-vacancies. For other tested strains, the unrelaxed-TDEs decreasewith respect to the amount of strains, same as TDEs but withlower rates. The unrelaxed-TDE and TDE in the static approachcould be considered as upper and lower bounds of damagethreshold energies in dynamic simulations respectively.Compared to the unrelaxed-TDE, as discussed previously, theTDE obtained from relaxed vacancy congurations in the staticapproach is better in characterizing the radiation hardness.

From the energy perspective, any effort leading to the incre-ment of system energy will potentially reduce the radiationhardness. This could also explain the temperature effect onradiation hardness.52 The total energy in a system increases as thetemperature increases, which could in turn reduce the radiationhardness. Introducing other defects will potentially increase thesystem energy, and thus minimize the radiation hardness. Asimilar explanation holds when electric53 and magnetic elds54

are applied. Currents and charges will introduce additional workto the system and increase the system's energy, thereby reducingthe radiation hardness. Such a general trend of radiation hard-ness could be widely applied to material design and engineeringfor those devices working in radiation-rich environments, e.g.,electronic and optoelectronic devices in outer space.

3.4 Radiation damage cross-sections

According to eqn (1), shd, the radiation damage cross-section forisotropic knock-on radiation with strain h can be obtainedthrough the calculation of threshold displacement energy Ehd,v.Fig. 5 shows the radiation damage cross-sections shd of producinga B monovacancy (le column) and a N monovacancy (rightcolumn) as a function of the electron beam energy and the strainh along the armchair, zigzag, and biaxial directions respectively.

For a xed incoming electron beam energy Ee, theminimum ofshd is at h¼ 0 for all six cases studied.With the same strain and thesame incoming electron beam energy Ee, the cross-section of

702 | Nanoscale, 2013, 5, 695–703

producing a N monovacancy is larger than that of producing a Bmonovacancy. This result is consistent with the change inthreshold displacement energy and electron energy. For B vacancycongurations with a strain h, shd has a maximum value withrespect to the electron beam energy Ee. However, for N vacancycongurations with a strain h, shd monotonically increases withrespect to the electron beam energy. The extreme points in shd � h

and shd� Ee curves indicate that radiation damage ismaximum forcertain strains and incoming electron energies. Thus, the contoursof shd could serve as a road map in engineering h-BN based deviceswhen the radiation resistance is the primary concern.

4 Conclusion

In summary, we studied the strain effect on the radiationhardness during controlled radiation damage of hexagonalboron nitride monolayers using ab initio density functionaltheory calculations. Three typical strain modes which are alongthe armchair, zigzag, and biaxial directions were applied. Twokinds of controlled radiation damage were studied which forma vacancy at a boron (VB) and a nitrogen (VN) atomic site,respectively. Radiation damage is seen to reduce Young'smodulus by 9% (5%) for vacancy at the B (N) atomic site. Thestrain eld is affected by the applied strains and the vacancyspecies. Large von Mises equivalent atomic strains are observedaround the vacancies. The distribution of rdiff, the difference inelectronic charge density between the systems with and withoutvacancies, also shows that the largest rdiff value is around thedefects. The distribution of rdiff was affected by the appliedstrains and the vacancy species, consistent with the distribu-tions of von Mises equivalent atomic strains.

The threshold displacement energy of h-BN monolayersubject to isotropic knock-on electron radiation was calculatedin three deformation modes for both compressive and tensilestrains. We observed a general reduction in radiation hardnessdue to strains along the zigzag, armchair, and biaxial directions.We provided an explanation based on energy arguments.

Radiation damage cross-sections for isotropic knock-onradiation damage by primary electrons were also studied. For axed incoming electron beam energy, the minimum shd is at h¼0 for all six cases studied. Under the same strain and the sameincoming electron beam energy Ee, the cross-section ofproducing a nitrogen monovacancy is larger than that ofproducing a boron monovacancy. For B vacancy congurationsunder a certain strain h, shd has amaximum value with respect tothe electron beam energy Ee. However, shd monotonicallyincreases with respect to the electron beam energy. Knowledgeof the cross-sections could serve as a road map in engineeringh-BN based devices under irradiation.

Acknowledgements

The authors would like to acknowledge the generous nancialsupport from the Defense Threat Reduction Agency (DTRA)Grant # BRBAA08-C-2-0130, the U.S. Nuclear RegulatoryCommission Faculty Development Program under contract #

This journal is ª The Royal Society of Chemistry 2013

Paper Nanoscale

NRC-38-08-950, and U.S. Department of Energy (DOE) NuclearEnergy University Program (NEUP) Grant # DE-NE0000325.

References

1 W.-Q. Han, L. Wu, Y. Zhu, K. Watanabe and T. Taniguchi,Appl. Phys. Lett., 2008, 93, 223103.

2 J. Kotakoski, C. H. Jin, O. Lehtinen, K. Suenaga andA. V. Krasheninnikov, Phys. Rev. B: Condens. Matter Mater.Phys., 2010, 82, 113404.

3 C. Jin, F. Lin, K. Suenaga and S. Iijima, Phys. Rev. Lett., 2009,102, 195505.

4 N. Alem, R. Erni, C. Kisielowski, M. D. Rossell, W. Gannettand A. Zettl, Phys. Rev. B: Condens. Matter Mater. Phys.,2009, 80, 155425.

5 J. C. Meyer, A. Chuvilin, G. Algara-Siller, J. Biskupek andU. Kaiser, Nano Lett., 2009, 9, 2683–2689.

6 O. L. Krivanek, M. F. Chisholm, V. Nicolosi, T. J. Pennycook,G. J. Corbin, N. Dellby, M. F. Murtt, C. S. Own, Z. S. Szilagyi,M. P. Oxley, S. T. Pantelides and S. J. Pennycook, Nature,2010, 464, 571–574.

7 Y. Ma, Y. Dai, W. Wei, C. Niu, L. Yu and B. Huang, J. Phys.Chem. C, 2011, 115, 20237–20241.

8 Z. H. Aitken and R. Huang, J. Appl. Phys., 2010, 107, 123531.9 G. Gui, J. Li and J. Zhong, Phys. Rev. B: Condens. Matter Mater.Phys., 2008, 78, 075435.

10 Z. H. Ni, T. Yu, Y. H. Lu, Y. Y. Wang, Y. P. Feng andZ. X. Shen, ACS Nano, 2008, 2, 2301–2305.

11 S.-M. Choi, S.-H. Jhi and Y.-W. Son, Nano Lett., 2010, 10,3486–3489.

12 Y. M. Lin, C. Dimitrakopoulos, K. A. Jenkins, D. B. Farmer,H. Y. Chiu, A. Grill and P. Avouris, Science, 2010, 327, 662.

13 Y. Ma, Y. Dai, M. Guo and B. Huang, Phys. Rev. B: Condens.Matter Mater. Phys., 2012, 85, 235448.

14 Y. Ma, Y. Dai, M. Guo, C. Niu, Y. Zhu and B. Huang, ACSNano, 2012, 6, 1695–1701.

15 Y. Ma, Y. Dai, M. Guo, C. Niu, J. Lu and B. Huang, Phys.Chem. Chem. Phys., 2011, 13, 15546–15553.

16 Q. Peng, A. R. Zamiri, W. Ji and S. De, Acta Mechanica, 2012,223, 2591–2596.

17 Q. Peng, W. Ji and S. De, Phys. Chem. Chem. Phys., 2012, 14,13385–13391.

18 A. Nag, K. Raidongia, K. P. S. S. Hembram, R. Datta,U. V. Waghmare and C. N. R. Rao, ACS Nano, 2010, 4, 1539.

19 Y.-N. Xu and W. Y. Ching, Phys. Rev. B: Condens. MatterMater. Phys., 1991, 44, 7787–7798.

20 N. Ooi, A. Rairkar, L. Lindsley and J. B. Adams, J. Phys.:Condens. Matter, 2006, 18, 97.

21 J. Furthmuller, J. Hafner and G. Kresse, Phys. Rev. B:Condens. Matter Mater. Phys., 1994, 50, 15606–15622.

22 Q. Tang, Z. Zhou and Z. Chen, J. Phys. Chem. C, 2011, 115,18531–18537.

23 Q. Peng and S. De, Physica E, 2012, 44, 1662–1666.24 E. Hernandez, C. Goze, P. Bernier and A. Rubio, Phys. Rev.

Lett., 1998, 80, 4502.25 A. Suryavanshi, M. Yu, J. Wen, C. Tang and Y. Bando, Appl.

Phys. Lett., 2004, 84, 2527.

This journal is ª The Royal Society of Chemistry 2013

26 P. Kim, L. Shi, A. Majumdar and P. L. McEuen, Phys. Rev.Lett., 2001, 87, 215502.

27 X. Blase, A. Rubio, S. G. Louie and M. L. Cohen, Europhys.Lett., 1994, 28, 335.

28 K. Watanabe, T. Taniguchi and H. Kanda, Nat. Mater., 2004,3, 404.

29 M. Hirayama and K. Shohno, J. Electrochem. Soc., 1975, 122,1671–1676.

30 G. W. Lee, M. Park, J. Kim, J. I. Lee and H. G. Yoon,Composites, Part A, 2006, 37, 727.

31 C. Zhi, Y. Bando, C. Tang, H. Kuwahara and D. Golberg, Adv.Mater., 2009, 21, 2889.

32 G. Giovannetti, P. A. Khomyakov, G. Brocks, P. J. Kelly andJ. van den Brink, Phys. Rev. B: Condens. Matter Mater. Phys.,2007, 76, 073103.

33 L. Ci, L. Song, C. Jin, D. Jariwala, D. Wu, Y. Li, A. Srivastava,Z. F. Wang, K. Storr, L. Balicas, F. Liu and P. M. Ajayan, Nat.Mater., 2010, 9, 430–435.

34 Q. Peng, W. Ji and S. De, Comput. Mater. Sci., 2012, 56,11–17.

35 J. S. Kim, K. B. Borisenko, V. Nicolosi and A. I. Kirkland, ACSNano, 2011, 5, 3977.

36 E. Holmstrom, L. Toikka, A. V. Krasheninnikov andK. Nordlund, Phys. Rev. B: Condens. Matter Mater. Phys.,2010, 82, 045420.

37 S. J. Zinkle and C. Kinoshita, J. Nucl. Mater., 1997, 251, 200.38 D. R. Locker and J. M. Meese, IEEE Trans. Nucl. Sci., 1972, 19,

237.39 J. V. Sharp and D. Rumsby, Radiat. Eff., 1973, 17, 65.40 F. Seitz and J. Koehler, Solid State Physics, Academic Press,

New York, 1956, vol. 2.41 A. Zobelli, A. Gloter, C. P. Ewels, G. Seifert and C. Colliex,

Phys. Rev. B: Condens. Matter Mater. Phys., 2007, 75, 245402.42 G. Kresse and J. Hafner, Phys. Rev. B: Condens. Matter Mater.

Phys., 1993, 47, 558.43 W. Kohn and L. J. Sham, Phys. Rev., 1965, 140, A1133.44 J. Perdew, K. Burke and M. Ernzerhof, Phys. Rev. Lett., 1996,

77, 3865.45 R. O. Jones and O. Gunnarsson, Rev. Mod. Phys., 1989, 61,

689–746.46 L. Liu, Y. P. Feng and Z. X. Shen, Phys. Rev. B: Condens. Matter

Mater. Phys., 2003, 68, 104102.47 C. F. von Weizsacker, Z. Phys., 1935, 96, 431.48 L. Song, L. Ci, H. Lu, P. B. Sorokin, C. Jin, J. Ni,

A. G. Kvashnin, D. G. Kvashnin, J. Lou, B. I. Yakobson andP. M. Ajayan, Nano Lett., 2010, 10, 3209.

49 C. Lee, X. Wei, J. W. Kysar and J. Hone, Science, 2008, 321,385.

50 F. Shimizu, S. Ogata and J. Li, Mater. Trans., 2007, 48, 2923–2927.

51 M. Utz, Q. Peng and M. Nandagopal, J. Polym. Sci., Part B:Polym. Phys., 2004, 42, 2057–2065.

52 P. M. Rice and S. J. Zinkle, J. Nucl. Mater., 1998, 258, 1414.53 V. Danchenk, U. D. Desai, J. M. Killiany and S. S. Brashear,

IEEE Trans. Electron Devices, 1968, ED15, 751.54 O. Cantoni, P. Sestili, M. Fiorani and M. Dacha, Biochem.

Mol. Biol. Int., 1996, 38, 527.

Nanoscale, 2013, 5, 695–703 | 703