Quantum mechanical modeling of hydrogen assisted cracking...

6
PHYSICAL REVIEW B 88, 104109 (2013) Quantum mechanical modeling of hydrogen assisted cracking in aluminum Yi Sun, 1 Qing Peng, 2 and Gang Lu 1 1 Department of Physics and Astronomy, California State University, Northridge, California, USA 2 Department of Mechanical, Aerospace, and Nuclear Engineering, Rensselaer Polytechnic Institute, Troy, New York, USA (Received 8 December 2012; revised manuscript received 5 September 2013; published 26 September 2013) We report multiscale quantum mechanical modeling of hydrogen assisted cracking in aluminum which is central to H embrittlement phenomena. We find that dislocation emission and brittle cleavage can occur simultaneously. H embrittlement takes place when H occupies the top sites on the crack front surface and even a very low H coverage at 0.2 monolayers can lead to brittle cleavage. H atoms adsorbed on the crack surfaces tend to suppress dislocation emission, whereas the solute H atoms on the slip plane can promote dislocation emission. Top-site H atoms at the front surface are found to facilitate the migration of other H atoms towards the front surface, providing a mechanism for H accumulation at the crack tip. The study resolves a long-standing puzzle of why H embrittlement could occur in Al where the equilibrium H solubility is extremely low under normal conditions. DOI: 10.1103/PhysRevB.88.104109 PACS number(s): 62.20.mt, 71.15.Mb, 81.40.Np I. INTRODUCTION Hydrogen embrittlement is one of the most pervasive problems in materials science and engineering as almost all metals and their alloys suffer to some extent of this H-induced brittleness. 1,2 H embrittlement has been responsible for many, if not most, service failures in materials applications where the components and structures come into contact with natural or technological environments including aqueous solution, gas, elevated temperature, irradiation, etc. However, despite enormous research efforts in past decades, 116 a complete mechanistic understanding of H embrittlement phenomena still eludes us. For the most part, the challenges stem from the multiscale nature of the problem—the interplay involving nanoscale processes (such as bond breaking and nucleation of defects) and long-range elastic interactions among the extended defects. Fortunately, thanks to ever increasing computational power and advances in multiscale modeling techniques 1724 pertinent atomistic simulations of H embrittlement began to emerge. 11,13,25 In this paper, we address two key issues at the core of H embrittlement: H assisted cracking and transport of H to the crack front by multiscale quantum mechanical simulations. The simulations illuminate the atomistic origin of H embrit- tlement, provide insight into the competing mechanisms of fracture, and offer a basis for comprehensive understanding of H embrittlement of metals. Owing to an extremely low H solubility in bulk Al (the atomic percentage in the range of 10 6 to 10 8 ), 26 H on the crack surfaces is found to play an important role in H embrittlement, which has been largely ignored in the past. These surface H atoms could originate from those trapped at vacancies or microvoids that eventually coalesce to form cracks. 6 We find that H assisted cracking occurs if H atoms occupy certain corrosive sites on the crack front surface. Here the existence of such corrosive sites is established and their role in H embrittlement is illustrated. We find that dislocation emission and brittle cleavage could happen simultaneously, challenging a traditional view that the two competing processes are mutually exclusive. In contrast to the conventional notion that H concentrations in metals are often too low to break the metallic bonds, 11 we observe that brittle fracture can occur even with a low H coverage (0.2 monolayers) at the corrosive sites. Moreover, the H atom at the corrosive site can facilitate the migration of interior H atoms towards the crack front surface for sustained H embrit- tlement. Therefore, the simulations resolve a long-standing puzzle of why H embrittlement could occur in Al where the bulk solubility of H is extremely low under normal conditions. Recently, an atomistic study has been conducted in Fe which reported a ductile to brittle transition caused by the suppression of dislocation emission at the crack tip due to aggregation of H. 13 In addition to the differences in computational methodologies, the present work focuses on the chemical aspect on the fracture surface (H assisted cracking) in Al. Since H solubility in Al is much lower than that in Fe, the H segregation mechanism observed in Ref. 13 may not be applicable to Al. II. METHODOLOGY Here, a semi-infinite crack in a single Al crystal under mode I loading is simulated using a quantum mechanical multiscale method: quasicontinuum density functional theory (QCDFT). 18,22 The QCDFT simulations are performed on a ductile crack tip orientation with ( ¯ 110) crack plane and [ ¯ 1 ¯ 12] crack front direction. The relevant directions and the schematic partition of simulation domains are shown in Fig. 1. In this orientation, the active slip planes for dislocation nucleation are (111) planes; all other {111} planes lie obliquely to the crack plane and are thus precluded from the simulations. 27 The dimensions of the system are 1.6 μm × 1.6 μm × 4.887 ˚ A along the [111] (x ), [ ¯ 110] (y ), [ ¯ 1 ¯ 12] (z) directions respectively. The system is periodic in the z direction and has Dirichlet boundary conditions in the other two directions. In the QCDFT simulations, the system is partitioned into three distinct domains: (1) a nonlocal quantum mechanical region (region I); (2) a nonlocal classical region (region II) where the classical EAM potential 28 is used; and (3) a local re- gion (region III) where the finite-element method is employed based on the Cauchy-Born rule and the same EAM potential as region II. The coupling between regions II and III is achieved via the quasicontinuum formulation, 17,2931 while the coupling between regions I and II is accomplished by the QM/MM scheme. 32,33 The number of nonlocal atoms and finite-elements 104109-1 1098-0121/2013/88(10)/104109(6) ©2013 American Physical Society

Transcript of Quantum mechanical modeling of hydrogen assisted cracking...

Page 1: Quantum mechanical modeling of hydrogen assisted cracking ...csun.edu/sites/default/files/prem_lu_10.pdf · PHYSICAL REVIEW B 88, 104109 (2013) Quantum mechanical modeling of hydrogen

PHYSICAL REVIEW B 88, 104109 (2013)

Quantum mechanical modeling of hydrogen assisted cracking in aluminum

Yi Sun,1 Qing Peng,2 and Gang Lu1

1Department of Physics and Astronomy, California State University, Northridge, California, USA2Department of Mechanical, Aerospace, and Nuclear Engineering, Rensselaer Polytechnic Institute, Troy, New York, USA

(Received 8 December 2012; revised manuscript received 5 September 2013; published 26 September 2013)

We report multiscale quantum mechanical modeling of hydrogen assisted cracking in aluminum which is centralto H embrittlement phenomena. We find that dislocation emission and brittle cleavage can occur simultaneously.H embrittlement takes place when H occupies the top sites on the crack front surface and even a very low Hcoverage at 0.2 monolayers can lead to brittle cleavage. H atoms adsorbed on the crack surfaces tend to suppressdislocation emission, whereas the solute H atoms on the slip plane can promote dislocation emission. Top-siteH atoms at the front surface are found to facilitate the migration of other H atoms towards the front surface,providing a mechanism for H accumulation at the crack tip. The study resolves a long-standing puzzle of why Hembrittlement could occur in Al where the equilibrium H solubility is extremely low under normal conditions.

DOI: 10.1103/PhysRevB.88.104109 PACS number(s): 62.20.mt, 71.15.Mb, 81.40.Np

I. INTRODUCTION

Hydrogen embrittlement is one of the most pervasiveproblems in materials science and engineering as almost allmetals and their alloys suffer to some extent of this H-inducedbrittleness.1,2 H embrittlement has been responsible for many,if not most, service failures in materials applications wherethe components and structures come into contact with naturalor technological environments including aqueous solution,gas, elevated temperature, irradiation, etc. However, despiteenormous research efforts in past decades,1–16 a completemechanistic understanding of H embrittlement phenomenastill eludes us. For the most part, the challenges stemfrom the multiscale nature of the problem—the interplayinvolving nanoscale processes (such as bond breaking andnucleation of defects) and long-range elastic interactionsamong the extended defects. Fortunately, thanks to everincreasing computational power and advances in multiscalemodeling techniques17–24 pertinent atomistic simulations of Hembrittlement began to emerge.11,13,25

In this paper, we address two key issues at the core of Hembrittlement: H assisted cracking and transport of H to thecrack front by multiscale quantum mechanical simulations.The simulations illuminate the atomistic origin of H embrit-tlement, provide insight into the competing mechanisms offracture, and offer a basis for comprehensive understandingof H embrittlement of metals. Owing to an extremely low Hsolubility in bulk Al (the atomic percentage in the range of10−6 to 10−8),26 H on the crack surfaces is found to play animportant role in H embrittlement, which has been largelyignored in the past. These surface H atoms could originatefrom those trapped at vacancies or microvoids that eventuallycoalesce to form cracks.6 We find that H assisted crackingoccurs if H atoms occupy certain corrosive sites on the crackfront surface. Here the existence of such corrosive sites isestablished and their role in H embrittlement is illustrated.We find that dislocation emission and brittle cleavage couldhappen simultaneously, challenging a traditional view that thetwo competing processes are mutually exclusive. In contrastto the conventional notion that H concentrations in metalsare often too low to break the metallic bonds,11 we observethat brittle fracture can occur even with a low H coverage

(0.2 monolayers) at the corrosive sites. Moreover, the H atomat the corrosive site can facilitate the migration of interior Hatoms towards the crack front surface for sustained H embrit-tlement. Therefore, the simulations resolve a long-standingpuzzle of why H embrittlement could occur in Al where thebulk solubility of H is extremely low under normal conditions.

Recently, an atomistic study has been conducted in Fewhich reported a ductile to brittle transition caused by thesuppression of dislocation emission at the crack tip dueto aggregation of H.13 In addition to the differences incomputational methodologies, the present work focuses on thechemical aspect on the fracture surface (H assisted cracking)in Al. Since H solubility in Al is much lower than that in Fe,the H segregation mechanism observed in Ref. 13 may not beapplicable to Al.

II. METHODOLOGY

Here, a semi-infinite crack in a single Al crystal undermode I loading is simulated using a quantum mechanicalmultiscale method: quasicontinuum density functional theory(QCDFT).18,22 The QCDFT simulations are performed on aductile crack tip orientation with (110) crack plane and [112]crack front direction. The relevant directions and the schematicpartition of simulation domains are shown in Fig. 1. In thisorientation, the active slip planes for dislocation nucleationare (111) planes; all other {111} planes lie obliquely to thecrack plane and are thus precluded from the simulations.27 Thedimensions of the system are 1.6 μm × 1.6 μm × 4.887 Aalong the [111] (x), [110] (y), [112] (z) directions respectively.The system is periodic in the z direction and has Dirichletboundary conditions in the other two directions.

In the QCDFT simulations, the system is partitioned intothree distinct domains: (1) a nonlocal quantum mechanicalregion (region I); (2) a nonlocal classical region (region II)where the classical EAM potential28 is used; and (3) a local re-gion (region III) where the finite-element method is employedbased on the Cauchy-Born rule and the same EAM potential asregion II. The coupling between regions II and III is achievedvia the quasicontinuum formulation,17,29–31 while the couplingbetween regions I and II is accomplished by the QM/MMscheme.32,33 The number of nonlocal atoms and finite-elements

104109-11098-0121/2013/88(10)/104109(6) ©2013 American Physical Society

Page 2: Quantum mechanical modeling of hydrogen assisted cracking ...csun.edu/sites/default/files/prem_lu_10.pdf · PHYSICAL REVIEW B 88, 104109 (2013) Quantum mechanical modeling of hydrogen

YI SUN, QING PENG, AND GANG LU PHYSICAL REVIEW B 88, 104109 (2013)

FIG. 1. (Color online) (a) An overview of the entire simulation box with finite-element mesh. (b) A blown-up view of the central boxin (a) showing the DFT region (red), EAM region (blue), and the finite-element region. (c) A displacement contour plot showing four edgedislocations that have been emitted and subsequently propagated 400 A away from the crack tip. All lengths are in A.

changes during the QCDFT simulation depending on the localdeformation gradients; a typical simulation includes about 450nonlocal DFT atoms, 500 nonlocal EAM atoms, and 5000finite elements, but overall, 82 million atoms are simulatedeffectively by QCDFT.

In region I, the quantum mechanical calculations are carriedout with the Vienna Ab initio Simulation Package (VASP)34,35

which is based on density functional theory (DFT) with thelocal density approximation and ultrasoft pseudopotentials.A plane-wave cutoff of 250 eV is used in the calculationsand the k points are sampled with a 1 × 1 × 11 mesh in theBrillouin zone.36 The EAM potential has been rescaled toreproduce a lattice constant and bulk modulus identical to thecorresponding VASP values.33

The simulations are carried out quasistatically with thedisplacement boundary conditions imposed at the exteriorcrack surfaces which are prescribed as a function of intendedstress intensity KI at each loading step. The loading procedureadopted here follows closely Ref. 22. The loading stepof 0.01 eV/A2.5 is used before the nucleation of the firstdislocation is observed. Once the crack tip plasticity istentatively detected, the load is backtracked and the loadingincrement is reduced to 0.001 eV/A2.5. This loading procedurehas been previously checked to produce converged results byusing the EAM-based QC method,22 which is computationallymuch more manageable.

The initial crack opening in the present work is determinedbased on two competing considerations: (1) It cannot be toonarrow; otherwise the crack will close and/or a large numberof loading steps would be required to observe the onset ofplasticity. (2) It cannot be too wide; otherwise the DFT regionwould be too large to render the calculations feasible. Ofcourse, the DFT region has to be large enough to capturethe crucial plasticity events at the crack tip. Based on the twoconsiderations, we have chosen the initial crack opening asfive layers in this work.

III. RESULTS AND ANALYSIS

H adsorptions on three crack surfaces are examined: theupper (U) and lower (L) side surfaces as well as the front (F)

crack surface as shown in Fig. 2(a). Several high-symmetryadsorption sites are considered on each of the surfaces,including face-centered cubic (fcc), hexagonal close-packed(hcp), top, and bridge sites. We find that the top and bridgesites are energetically more stable than the fcc and hcp sitesunder tensile strains. Hence in the following, we focus on thetop sites labeled by T and the bridge sites labeled by B in thecrack simulations.

As shown in Fig. 3, we have determined H adsorptionenergy on the (111) crack surface as a function of interplanarseparation distance δ. For each δ, the H adsorption energy isdefined as Ead(δ) = EAl+H(δ) − EAl(δ) − 1

2EH2 , where EAl+H

and EAl are the cohesive energies of the Al slab with andwithout a H atom and EH2 is the cohesive energy of an isolatedH2 molecule. The calculation is performed by using a slab with36 Al atoms via VASP. The dimensions of the slab are 8.5 A,4.9 A, 15.0 A along the [110], [112], and [111] direction,respectively. Vacuum is placed in the [111] direction on bothsides of the slab, and the periodic boundary conditions are

FIG. 2. (Color online) (a) Schematic diagram showing the threecrack surfaces: the upper side surfaces (U), the lower side surfaces(L), and the crack tip front surface (F). The slip plane positionsare indicated by xU in the upper half crystal and xL in the lower halfcrystal relative to the crack front. (b) Graphical definition of the crackopening width � = dy

′ − dy , where dy′ is the maximum interplanar

distance in the y direction and dy = 1.41 A is the equilibriuminterplanar distance.

104109-2

Page 3: Quantum mechanical modeling of hydrogen assisted cracking ...csun.edu/sites/default/files/prem_lu_10.pdf · PHYSICAL REVIEW B 88, 104109 (2013) Quantum mechanical modeling of hydrogen

QUANTUM MECHANICAL MODELING OF HYDROGEN . . . PHYSICAL REVIEW B 88, 104109 (2013)

FIG. 3. (Color online) The atomic configurations used in thecalculations of H adsorption energy Ead: (a) for a H atom on the B siteand (b) for a H atom on the T site on Al (111) surface. (c) Partiallyrelaxed H adsorption energy Ead versus the interplanar separation δ,showing the crossover between the B site and the T site. (d) Fullyrelaxed Ead versus δ. When δ reaches 0.9 A, the H atom shifts fromthe B site to the T site, accompanied by an energy jump. The red andblue spheres represent H and Al atoms, respectively.

applied in all directions. The k points are sampled accordingto the Monkhorst-Pack scheme with a 11 × 21 × 1 mesh in theBrillouin zone. The force convergence criterion is 0.02 eV/A.We find that at equilibrium (δ = 0) H is energetically morestable on the B site (Ead = 0.17 eV) than on the T site(Ead = 0.28 eV). When the crack opens up, the adsorptionenergy on the B site increases rapidly, in contrast to the case onthe T site, which remains approximately the same as shown inFig. 3(c). At δ = 1.2 A, the T site becomes energetically morestable than the B site. Note that the energy curves in Fig. 3(c)are calculated by fixing Al atoms (allowing H to move in the[111] direction) so that we can estimate the critical δ at whichthe energy crossover occurs. In Fig. 3(d), we present the resultswith a full relaxation of all atoms; we find that the B-site Hatom would shift to T site when δ reaches 0.9 A by breakingone of the two Al-H bonds. Therefore, for a typical crack tipconfiguration, the tension at the crack front can stabilize the Tsites and render them more favorable.

We have performed QCDFT calculations for ten distinctcrack configurations, including the pure Al crack as referenceand nine other cracks with different numbers of H atoms at theT sites and B sites on the three crack surfaces. We distinguishthese cracks by numbers from i to ix and summarize their mainfeatures in Table I. The number and position of H atoms areused to label the cracks in the second column of the Table.

For example, for crack ix, there are two T-site H atoms on theupper surface (U), two T-site H atoms on the lower surface(L), and five T-site H atoms on the front surface (F), indicatedby FT

5 +UT2 +LT

2 . As shown in Fig. 2, the fracture behavior ischaracterized by three quantities: (1) the critical stress intensityKIC under which the first full edge dislocation is emitted fromthe crack tip; (2) the position of the dislocation slip planes inthe upper (xU) and lower half (xL) of the crack plane. Thepositions xU and xL are given in terms of the number of (111)layers. A positive (negative) number signifies that the slip planeis on the right (left) side of the crack front surface. The slipplane is indicated by the sheared finite-element mesh. (3) Thecrack opening �, which is defined as � = dy

′ − dy , wheredy

′ is the maximum interplanar distance in the y directionand dy = 1.41 A is the equilibrium interplanar distance. � is

evaluated at KI = 0.30 eV/A2.5

for all cracks. � representsthe maximum lattice expansion between two adjacent atomicplanes at the crack tip, which is different from the crack-tip-opening displacement that measures the overall elastic openingof the crack tip.

It is found that for pure Al, KIC = 0.28 eV/A2.5 and thecorresponding value for the first partial dislocation is slightlylower, between 0.27 and 0.28 eV/A2.5. For all cracks examinedhere, it is much easier to identify the nucleation of a fulldislocation than a partial dislocation because the full edgedislocation is nucleated away from the crack tip, hence clearlyidentifiable in the displacement contour plot; one such exampleis shown in Fig. 1(c). For this reason, in Table I we onlylist KIC for the nucleation of the first full dislocation(s).Our KIC value is lower than that reported in Ref. 23 withKIC = 0.30 eV/A2.5, but higher than that reported in Ref. 25with KIC = 0.25 eV/A2.5. This is because a narrower cracktends to have a lower KIC, and the crack width in the presentstudy (5 layers, KIC = 0.28 eV/A2.5) is narrower than that inthe first report (6 layers, KIC = 0.30 eV/A2.5) and wider thanthat in the second report (3 layers, KIC = 0.25 eV/A2.5). Asshown in Fig. 4(a), increasing KI to 0.32 eV/A2.5, two moreedge dislocations are nucleated, one on each side of the crackplane, but without cleavage (� = 0.16 A). For the maximumKI = 0.36 eV/A2.5 applied in this study, the crack remainsductile. When H is adsorbed on the upper and lower surfaces(cracks i and ii), it has very little effect on the fracture behavior

TABLE I. Summary of KIC, �, and xU(L) for the ten cracks. NH

is the total number of adsorbed H atoms on the three crack surfaces.The figure references for the cracks are also included.

KIC � FigureCrack Configuration NH (eV/A2.5) xU xL (A) Reference

pure pure 0 0.28 −1/2 −1/2 0.16 Fig. 4(a)i UB

2 + LB2 4 0.29 −1/2 −1/2 1.02 Fig. 4(b)

ii UT2 + LT

2 4 0.28 −1/2 −1/2 0.21 Fig. 4(c)iii FB

6 6 0.32 −1/2 −1/2 1.21 Fig. 4(e)iv FT

1 1 0.31 1/2 1/2 6.16v FT

5 5 0.32 5/2 1/2 9.07 Fig. 4(f)vi FB

6 + UB2 8 0.30 −3/2 −1/2 1.35

vii FB6 + UB

2 + LB2 10 0.32 −3/2 −3/2 1.43

viii FT5 + UT

2 + LT2 9 0.30 1/2 1/2 8.42 Fig. 5(a)

ix H@slip plane 6 0.25 −1/2 −1/2 0.22 Fig. 4(d)

104109-3

Page 4: Quantum mechanical modeling of hydrogen assisted cracking ...csun.edu/sites/default/files/prem_lu_10.pdf · PHYSICAL REVIEW B 88, 104109 (2013) Quantum mechanical modeling of hydrogen

YI SUN, QING PENG, AND GANG LU PHYSICAL REVIEW B 88, 104109 (2013)

FIG. 4. (Color online) Schematic diagrams showing dislocation emission and brittle cleavage under KI = 0.32 eV/A2.5. Red and bluecircles represent H and Al atoms, respectively. All lengths are in A. The sheared finite-element mesh is the result of dislocation(s) passagealong the slip plane; more emitted dislocations yield more severe shears. (a) Pure Al; (d) solute H atoms at the slip plane (crack ix); (b) and(e) H atoms at the B sites of the side surfaces (crack i) and the front surface (crack iii), respectively; (c) and (f) H atoms at the T sites of theside surfaces (crack ii) and the front surface (crack v), respectively. The bond rupture in (f) is represented by the dashed lines.

of the cracks as displayed in Figs. 4(b) and 4(c). KIC for cracksi and ii remains approximately the same as in the pure crack,and no cleavage is observed even as KI is ramped up to 0.36eV/A2.5; instead, more dislocations are emitted from the cracktip. Therefore, the H atoms adsorbed on the side surfaces arenot responsible for H embrittlement of Al. However, whenH is adsorbed on the crack front surface, both ductile andbrittle behavior can occur depending on the adsorption sites.In Fig. 4(e), we illustrate the ductile response of the crack withsix H atoms occupying the B sites on the crack front (crackiii). Multiple dislocations are emitted but with no indicationof cleavage (� = 1.21 A) even as KI is increased to 0.36eV/A2.5. On the other hand, brittle fracture takes place whenH is adsorbed on the T site of the crack front (cracks iv and v).As depicted in Fig. 4(f), several Al-Al bonds are broken and thecrack grows by one atomic layer (� = 9.07 A). Interestingly,dislocations are also emitted simultaneously. The critical stressintensity KIC of 0.32 eV/A2.5 is higher than the correspondingvalue for the pure crack. One of the slip planes shifts forwardto xU = 5/2 as the result of crack propagation. Remarkably,the H assisted cracking is observed even with a single H atomat the T site of the crack front (crack iv); this corresponds toH coverage of 0.2 monolayers. In contrast, if H is adsorbedat the B site of the crack front, brittle cleavage cannot occureven at a much higher H coverage. For example, the cracks viand vii do not exhibit brittle behavior although there are manymore (8 and 10) H atoms on the crack surfaces. The increase

of KIC suggests that the surface-adsorbed H can suppressdislocation emission, which agrees with previous simulationresults.25 On the other hand, the solute H atoms at the slipplane (crack ix) can actually promote dislocation nucleationwith KIC reducing to 0.25 eV/A2.5, in line with the H enhancedlocal plasticity model1. This observation is consistent with theexperimental findings that when solute H was introduced inAl, dislocation activity was significantly enhanced ahead ofthe crack tip.4,5,37–39 We note that the quantitative results inTable I depend on H coverage at the crack tip; i.e., if the crackwidth and/or length changes, the quantitative results wouldchange accordingly.

We next pay closer attention to the brittle cleavage. InFig. 5(a), we display the brittle cleavage with H atoms onthe T sites of the front surface with KI = 0.30 eV/A2.5 (crackviii). If H migrates from the side surfaces to occupy the T sitesof the freshly exposed crack front as shown in Fig. 5(b), thebrittle cleavage can propagate one layer further. Conversely,if additional H cannot reach the fresh crack front, the brittlefracture halts as displayed in Fig. 5(c). Similarly, if H arrivesat the B site of the fresh crack front, the cleavage cannotgrow either as shown in Fig. 5(d). Therefore, H embrittlementproceeds by the corrosive H atoms attacking Al-Al bonds atthe front surface, one layer at a time. These T sites are hencethe active or corrosive sites for H embrittlement of Al.

The T site is corrosive because the formation of H-Al bondsat the T site significantly weakens Al-Al bonds on the crack

104109-4

Page 5: Quantum mechanical modeling of hydrogen assisted cracking ...csun.edu/sites/default/files/prem_lu_10.pdf · PHYSICAL REVIEW B 88, 104109 (2013) Quantum mechanical modeling of hydrogen

QUANTUM MECHANICAL MODELING OF HYDROGEN . . . PHYSICAL REVIEW B 88, 104109 (2013)

FIG. 5. (Color online) Schematic diagrams showing dislocationemission and brittle cleavage. (a) The initial cleavage for crack viiiunder KI = 0.30 eV/A2.5 (crack viii). (b) The crack extends onelayer forward with H at the T sites of the fresh front surface underKI = 0.35 eV/A2.5. (c) The crack halts without H on the fresh frontsurface. (d) The crack halts with H at the B sites of the fresh frontsurface. The red and purple dashed lines indicate the bond rupture onthe initial and new crack front, respectively.

front as H draws electrons from Al. When the weakened Al-Albonds are stretched at the crack tip, they can break easily. Onthe other hand, the formation of Al-H-Al bonds at the B sitecan actually stabilize the Al-Al bond on the crack front withH serving as a “glue” or linker. To provide direct evidence forthe above statement, we have calculated site-projected localdensity of states (LDOS) for the Al-Al bond in question usingthe model shown in Fig. 3 with δ = 0. The Al-Al bond isindicated by the red dashed line. Here we introduce a quantity� defined as � = ∫ Ef

−∞LDOSI (E) × LDOSII (E)dE, whereEf is the Fermi energy and LDOSI (E), LDOSII (E) are theLDOS of the two bonded Al atoms. � therefore represents“bond order” between the two Al atoms, which is proportionalto the overlapping density of states. We have calculated � forthree cases: Al-Al bond without H (�0), Al-Al bond with aT-site H (�T), and Al-Al bond with a B-site H (�B). We findthat �B = 0.657, �0 = 0.598, and �T = 0.590 in a rankingthat supports our statement.

Given the importance of the corrosive sites, the migrationenergy barriers of H from the bulk interior to the corrosivesites on the crack front are expected to affect the rate ofH embrittlement. There are two possible H sources inside amaterial: H adsorbed on the internal crack surfaces and soluteH in the bulk lattice. We find that the H diffusion barrier on theside surface and the front surface is 0.12 eV and 0.16 eV,respectively, lower than that in the bulk lattice, which is0.18 eV.26,40 However, it is also found that the corners betweenthe side surfaces and the front surface could act as bottlenecksfor H migration to the crack front and the correspondingmigration barrier is 0.24 eV. Therefore, the transport fromthe side surfaces to the front surface is not a preferential pathfor H arriving on the front surface. We have also examined

FIG. 6. (Color online) (a) The atomic structure of the crack tipused in the NEB calculations with one H atom at the T site of the frontsurface and another H migration towards the front surface. The arrowindicates the approximate initial and final positions of the migratingH atom. (b) H migration energy barriers from the bulk interior to thecrack front surface with and without a H atom at the T site. Eb is theH binding energy as defined in Ref. 41. The first Al layer is taken asthe front surface.

H diffusion barriers from the bulk interior to the crack frontalong the direction shown schematically in Fig. 6(a). A DFTcluster of 40.0 A × 35.0 A × 9.74 A carved out from therelaxed QCDFT atomic structure is used in the calculations,with three outmost Al layers fixed and other atoms allowed torelax.

We place one H atom at the T site of the front surfaceand load the crack using QCDFT until KI = 0.27 eV/A2.5.Under this loading, the crack is significantly deformed withthe crack opening � = 1.84 A at the front surface, butremains stable and dislocation free. The energy barriers foranother H along the migration path are calculated using thenudged elastic band (NEB) method.42 For comparison, we alsocalculate the corresponding energy barriers in the absence of Hat the front surface and the results are displayed in Fig. 6(b).The local minima and maxima in Fig. 6(b) represent the Hbinding energies at various distorted tetrahedral and octahedralinterstitial sites along the diffusion path. In the absence of Hat the front surface, the diffusion barriers are similar to that inthe bulk lattice. But with H at the front surface, the diffusionbarriers are drastically different; they decrease gradually from0.19 eV in the interior to 0.05 eV near the surface andcompletely vanish at the subsurface. Some of the octahedralsites are deformed so much that they resemble the tetrahedralsites, hence the diminished energy barriers. Therefore, theH atom at the front surface can facilitate the migration ofother H atoms to the front surface or the crack tip, whichprovides a mechanism for H multiplication and aggregation at

104109-5

Page 6: Quantum mechanical modeling of hydrogen assisted cracking ...csun.edu/sites/default/files/prem_lu_10.pdf · PHYSICAL REVIEW B 88, 104109 (2013) Quantum mechanical modeling of hydrogen

YI SUN, QING PENG, AND GANG LU PHYSICAL REVIEW B 88, 104109 (2013)

the crack front. The H atoms at the corrosive sites not only canbreak Al-Al bonds; they can also induce considerable latticestretching near the crack tip to help “pull” other H atoms out(the corresponding lattice stretching is much smaller withoutH at the front surface).

IV. CONCLUSION

In summary, we have studied H assisted cracking and Hdiffusion in Al cracks using multiscale quantum mechanicalsimulations. We find that dislocation emission and brittlecleavage could occur simultaneously. H embrittlement occurswhen H occupies the T sites on the crack front and evena low H coverage at 0.2 monolayers can lead to brittlecleavage. The H atoms on the crack surfaces tend to inhibitdislocation emission, whereas the solute H atoms at the slipplane promote dislocation emission. H at the crack front isfound to facilitate the diffusion of other H atoms to the

front surface, providing a mechanism for H aggregation atthe crack front. The simulations explain why a sufficientlyhigh H concentration at the crack tip, which is difficult torealize in Al under normal conditions, is in fact unnecessaryfor H embrittlement, resolving a long-standing puzzle in Hembrittlement of Al. The position of H is found to be moreimportant than its concentration for embrittlement of Al.Therefore H embrittlement could be mitigated if the T sites areblocked or made less stable by alloying, heat treatment, or othermeans. The understanding and control of crack tip chemistryholds the key to rational design of embrittlement-resistantmaterials for which multiscale quantum mechanical modelingcould be an invaluable tool.

ACKNOWLEDGMENT

We gratefully acknowledge the support from the Office ofNaval Research.

1S. M. Myers et al., Rev. Mod. Phys. 64, 559 (1992).2N. R. Moody and A. W. Thompson, in Hydrogen Effects on MaterialBehavior (TMS AIME, Warrendale, PA, 1990).

3S. P. Lynch, in Gaseous Hydrogen Embrittlement of Materials inEnergy Technologies (Woodhead, Cambridge, 2012).

4G. M. Bond, I. M. Robertson, and H. K. Birnbaum, Acta Metall.35, 2289 (1987).

5G. M. Bond, I. M. Robertson, and H. K. Birnbaum, Acta Metall.36, 2193 (1988).

6G. Lu and E. Kaxiras, Phys. Rev. Lett. 94, 155501 (2005).7G. Lu, Q. Zhang, N. Kioussis, and E. Kaxiras, Phys. Rev. Lett. 87,095501 (2001).

8Y. Liang, P. Sofronis, and N. Aravas, Acta Mater. 51, 2717 (2003).9J. Song, M. Soare, and W. A. Curtin, Modell. Simul. Mater. Sci.Eng. 18, 045003 (2010).

10D. E. Jiang and E. A. Carter, Acta Mater. 52, 4801 (2004).11R. Matsumoto, Y. Inoue, S. Taketomi, and N. Miyazaki, Scr. Mater.

60, 555 (2009).12J. Song and W. A. Curtin, Acta Mater. 59, 1557 (2011).13J. Song and W. A. Curtin, Nat. Mater. 12, 145 (2013).14R. Matsumoto, S. Taketomi, S. Matsumoto, and N. Miyazaki, Int.

J. Hydrogen Energy 34, 9576 (2009).15S. Taketomi, R. Matsumoto, and N. Miyazaki, J. Mater. Sci. 43,

1166 (2008).16K. Takayama, R. Matsumoto, S. Taketomi, and N. Miyazaki, Int. J.

Hydrogen Energy 36, 1037 (2011).17G. Lu and E. Kaxiras, in Theoretical and Computational Nan-

otechnology (American Scientific Publisher, Stevenson Ranch, CA,2005), Chap. 22.

18G. Lu, E. B. Tadmor, and E. Kaxiras, Phys. Rev. B 73, 024108(2006).

19Q. Peng, X. Zhang, L. Hung, E. A. Carter, and G. Lu, Phys. Rev. B78, 054118 (2008).

20A. K. Nair, D. H. Warner, R. G. Hennig, and W. A. Curtin, Scr.Mater. 63, 1212 (2010).

21S. A. Serebrinsky, E. A. Carter, and M. Ortiz, J. Mech. Phys. Solids52, 2403 (2004).

22Q. Peng and G. Lu, J. Mech. Phys. Solids 59, 775 (2011).

23L. Hung and E. A. Carter, Modelling Simul. Mater. Sci. Eng. 19,045002 (2011).

24S. Ogata, F. Shimojo, R. K. Kalia, A. Nakano, and P. Vashishta,J. Appl. Phys. 95, 5316 (2004).

25R. J. Zamora, A. K. Nair, R. G. Hennig, and D. H. Warner, Phys.Rev. B 86, 060101 (2012).

26C. Wolverton, V. Ozolins, and M. Asta, Phys. Rev. B 69, 144109(2004).

27E. B. Tadmor and S. Hai, J. Mech. Phys. Solids 51, 765 (2003).28F. Ercolessi and J. B. Adams, Europhys. Lett. 26, 583 (1994).29V. B. Shenoy, R. Miller, E. B. Tadmor, D. Rodney, R. Phillips, and

M. Ortiz, J. Mech. Phys. Solids 47, 611 (1999).30E. B. Tadmor, R. Miller, and R. Phillips, J. Mater. Res. 14, 2249

(1999).31R. Miller, E. B. Tadmor, R. Phillips, and M. Ortiz, Modelling Simul.

Mater. Sci. Eng. 6, 607 (1998).32Y. Liu, G. Lu, Z. Chen, and N. Kioussis, Simul. Mater. Sci. Eng.

15, 275 (2007).33N. Choly, G. Lu, W. E, and E. Kaxiras, Phys. Rev. B 71, 094101

(2005).34G. Kresse and J. Hafner, Phys. Rev. B 47, 558 (1993).35G. Kresse and J. Hafner, Phys. Rev. B 49, 14251 (1994).36H. J. Monkhorst and J. D. Pack, Phys. Rev. B 13, 5188 (1976).37P. J. Ferreira, I. M. Robertson, and H. K. Birnbaum, Acta Mater.

47, 2991 (1999).38P. J. Ferreira, I. M. Robertson, and H. K. Birnbaum, Acta Mater.

46, 1749 (1998).39P. J. Ferreira, I. M. Robertson, and H. K. Birnbaum, Materials

Science Forum 207-209, 93 (1996).40G. Lu, D. Orlikowski, I. Park, O. Politano, and E. Kaxiras, Phys.

Rev. B 65, 064102 (2002).41Binding energy Eb computed as Eb = Ecluster+H − Ecluster − 1

2 EH2 .Ecluster+H and Ecluster are the total energies of the cluster with andwithout the diffusing H atom, respectively, and EH2 is the totalenergy of an isolated H2 molecule.

42H. Jonsson, G. Mills, and K. W. Jacobsen, in Classical and QuantumDynamics in Condensed Phase Simulations, (World Scientific,Singapore, 1998), p. 385.

104109-6