Download as pdf or txt
Download as pdf or txt
You are on page 1of 24

SAND2016-1481J

Molecular Dynamics Simulations of Hydrogen Diffusion in Aluminum


X. W. Zhou*, F. El Gabaly, V. Stavila, and M. D. Allendorf

Sandia National Laboratories, Livermore, California 94550, USA

ABSTRACT

Hydrogen diffusion impacts the performance of solid-state hydrogen storage materials, and

contributes to the embrittlement of structural materials under hydrogen-containing environments.

In atomistic simulations, the diffusion energy barriers are usually calculated using molecular

statics simulations where a nudged elastic band method is used to constrain a path connecting the

two end points of an atomic jump. This approach requires prior knowledge of the “end points”.

For alloy and defective systems, the number of possible atomic jumps with respect to local

atomic configurations is tremendous. Even when these jumps can be exhaustively studied, it is

still unclear how they can be combined to give an overall diffusion behaviour seen in

experiments. Here we describe the use of molecular dynamics simulations to determine the

overall diffusion energy barrier from the Arrhenius equation. This method does not require

information about atomic jumps, and it has additional advantages, such as the ability to

incorporate finite temperature effects and to determine the pre-exponential factor. As a test case

for a generic method, we focus on hydrogen diffusion in bulk aluminium. We find that the

challenge of this method is the statistical variation of the results. However, highly converged

energy barriers can be achieved by an appropriate set of temperatures, output time intervals (for

tracking hydrogen positions), and a long total simulation time. Our results help elucidate the

inconsistencies of the experimental diffusion data published in literature. The robust approach

*
xzhou@sandia.gov

1
developed here may also open up future molecular dynamics simulations to rapidly study

diffusion properties of complex material systems in multi-dimensional spaces involving

composition and defects.

Keywords: molecular dynamics, hydrogen diffusion, aluminium

I. INTRODUCTION

As an important rate-limiting mechanism for kinetic processes in solids, mass diffusion is

often the barrier to the development of many new material technologies. For example, replacing

fossil fuels with carbon-free hydrogen for transportation energy is widely explored. Some solid-

state hydrogen storage materials can provide sufficient hydrogen for vehicles [1,2]. However,

many of the high-capacity materials suffer from impractical hydrogen charging and discharging

rates, thought to be due in part to the slow diffusion kinetics [2,3]. To develop storage materials

that can meet DOE target for refuelling time, the fundamental mechanisms and energy barriers of

diffusion in hydrogen storage materials must be understood [2]. The challenge is that the

hydrogen charging / recharging cycles are associated with significant changes in both

compositions and structures (e.g., phase transformation, interface migration, and defect

formation). This material complexity requires studying diffusion in evolving structures, which

has inhibited developing new hydrogen storage materials [4].

In another example, structural materials under hydrogen-containing environments are

needed to store and deliver hydrogen. Hydrogen embrittlement is the limiting factor for these

materials [5]. Hydrogen diffusion to defects (e.g., grain boundaries, dislocations, crack tips, etc.),

directly contributes to the deterioration of mechanical properties with time. Understanding

hydrogen diffusion is equally important to develop new structural materials.

2
Atomistic simulations such as molecular dynamics (MD) and molecular statics (MS)

provide a theoretical means to study diffusion. While the accuracy of results depends on the

interatomic potential used in the simulations, the advantages of these methods are that they do

not require assumptions, enable complex atomic scale structures to be accurately replicated in the

computational cell, and allow diffusion mechanisms (e.g., atom jumps) to be visualized. In the

past, diffusion energy barriers were usually calculated using molecular statics (i.e., energy

minimization) rather than molecular dynamics simulations. By using a nudged elastic band

method [6,7,8] to constrain a path connecting the two end points of an atomic jump, a single MS

simulation can determine the entire minimum energy path and the associated energy barrier. This

approach requires prior knowledge of the “end points” of atomic jumps. In addition, the number

of possible atomic jumps with respect to local atomic configurations is tremendous for alloy and

defective systems. Even if these diffusion paths can be exhaustively studied, it is still unclear

how they can be combined to give an overall diffusion behaviour seen in experiments.

Alternatively, molecular dynamics simulations can be used to track the mean square

displacement of diffusion atoms over a given period of time at different temperatures. The

overall diffusion energy barrier can then be determined from the Arrhenius equation regardless

of the number of the atomic jumps. This method has additional advantages such as incorporating

finite temperature effects and determining the pre-exponential factor. Unfortunately, this method

often has unsatisfactorily large statistical errors, and hence has not been widely used in the past.

Using the diffusion of a single hydrogen atom in bulk aluminium as a test case, the

objective of the present work is to both develop molecular dynamics models capable of

accurately quantifying diffusion energy barriers from the Arrhenius equation, and understand

hydrogen diffusion in aluminium. The MD code LAMMPS [9,10] is used for this study. To

3
provide researchers with an useful handbook reference missing in literature, we will examine

various ensembles including NVT (constant number of atoms, volume, and temperature), NVE

(constant number of atoms, volume, and energy), and NPT (constant number of atoms, pressure,

and temperature). Both Nose-Hoover [11] and Langevin [12] algorithms are used to control

temperature and pressure. Effects of various MD parameters (e.g., damping and dragging

coefficients of thermostat and barostat, etc.) implemented in LAMMPS are all explored. We will

validate the calculations using experimental data available in literature. The Al-H system is ideal

for the study because the hydrogen diffusion in aluminium is known to be sensitive to

composition and structure [13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31] and it is

this complexity that needs to be understood for hydrogen storage materials. Theoretical studies

of this system also help understand the inconsistent experimental diffusion measurements [13,14,

15,16,17,18,19,20,21,22,23,24,25,26,27,28,29] made on Al-H. Furthermore, aluminum is a light

weight structural material suitable for vehicle applications [32]. In particular, no hydrogen

embrittlement has been reported for aluminum alloys when they are not used under fatigue or

corrosive conditions [33,34,35]. This is in sharp contrast with steels [36]. As a result,

understanding hydrogen diffusion in aluminum can have a high impact on a range of applications

including use of this metal as a structural material.

Our work will be based on an Al-H bond order potential (BOP) [37,38], which we have

shown to capture energy and volume trends for a variety of Al, H, and Al-H structures that are

the determining factors for hydrogen diffusion. It also captures some important but elusive

properties such as the high stacking fault energy of aluminium [37], the energy barrier of the H +

H2  H + H exchange reaction [38], and the Al-H phase diagram [37].

II. MOLECULAR DYNAMICS METHODS

4
Our single crystalline, face-centred-cubic (fcc) aluminium cells contain 8 {100} planes in

each of the three <100> coordinate directions. The initial crystals are created based on the lattice

constant derived from the interatomic potential [37], a = 4.0494 Å. The system dimension is

therefore around 32 Å3, corresponding to 2048 aluminium atoms. Bulk crystals are simulated

using periodic boundary conditions, and a single hydrogen atom is introduced in each simulated

computational cell. The target temperature T is created by assigning velocities to atoms based on

the Boltzmann distribution, which is further equilibrated by performing a warm-up MD

simulation for more than 0.1 ns. After this, MD simulations of hydrogen diffusion are performed

for a total period of tMD, which corresponds to a total time steps of nMD = tMD/dt where dt is the

time step size.

During MD simulations, the hydrogen location as a function of time, r(t), is tracked on a

time interval of t, i.e., at times of t = it, i = 1, 2, …, m (m = tMD/t), where t is a multiple of

dt. The relative hydrogen displacement per t, measured between (i-1)t and it, is then

calculated as ri(t) = r(it) - r(it - t), i = 1, 2, …, m. Once r per t is known, the relative

displacement per larger time intervals of kt, measured between (i-1)t and (i-1+k)t, can be

i 1 k
simply obtained as ri(kt) =  r t  . Clearly, for a given kt, we can compute m+1-k values
j i
j

of ri(kt): i = 1, 2, …, m+1-k. As a result, the number of ri(kt) values reduces to one when

k approaches m, but there are many ri(kt) values when k << m. Therefore, under the condition

k << m, we can calculate highly converged mean square displacement from:

m 1 k

 r kt 
2
i
S kt 
2 i 1
 (1)
m 1 k

5
The mean square displacement S t  2 is a linear function of time t. In particular, S t  2 = 6Dt

where D is diffusivity [39]. Fitting the MD S t  2 data to 6Dt in a small time range t << tMD

(i.e., k << m) allows us to determine diffusivity D.

The MD procedures described above can be used to calculate diffusivity at different

temperatures. The results can then be fitted to the Arrhenius equation to get the activation energy

of hydrogen diffusion, Q, and the pre-exponential factor, D0.

III. RESULTS

A. Molecular Statics Calculations

For comparison, we first calculate hydrogen diffusion energy barrier using MS simulations

in conjunction with a nudged elastic band method, based on the same interatomic potential. The

diffusion of a single hydrogen atom in an aluminium bulk is a special cause where MS

simulations can be applied. This is because there are only four possible diffusion paths:

octahedral to octahedral (O-O), tetrahedral to tetrahedral (T-T), octahedral to tetrahedral (O-T),

and tetrahedral to octahedral (T-O) jumps. We find that the O-O and T-T jumps have higher

energy barriers and hence diffusion always occurs through the O-T and T-O jumps. Furthermore,

our calculations show an activation energy of Q  0.41 eV for the O-T jump, and Q  0.37 eV

for the T-O jump. For continuous diffusion alternating between T-O and O-T jumps, the higher

value of Q  0.41 eV is taken as the overall activation energy barrier.

B. Convergence of MD Results

Highly converged activation energies require long MD simulations. Somewhat surprisingly,

we also find that the output time interval t, and the simulation temperatures T, impact the

convergence. Hence, we first discuss the t and T effects.

6
Based on an NVT ensemble, two sets of MD simulations are performed using a total MD

time of tMD = 0.88 ns and a time step size of dt = 0.001 ps (i.e., a total MD steps of nMD =

8.8105). The first set uses t = 0.0088 ps and various temperatures between 400 and 800 K (if

normalized by the activation energy Q = 0.41 eV as determined from MS simulations above, this

temperature range corresponds to T/Q = 976 – 1951 K/eV†). The second set uses t = 4.4 ps and

various temperatures between 500 and 700 K (T/Q = 1220 – 1707 K/eV). The mean square

displacement of the hydrogen atom is calculated as a function of time according to Eq. (1), and

the results are fitted to 6Dt for a small time range (t < 15 ps). A small time range is necessary to

produce sufficient terms in the average sum. Specifically, the total number of terms in Eq. (1)

can be written as N = m + 1 – k = tMD/t + 1 – t/t. Hence, for our chosen time range t < 15 ps, N

> 98296 at t = 0.0088 ps, and N > 1966 at t = 4.4 ps. The MD mean square displacement

results and the fitted 6Dt lines are shown in Fig. 1(a) for three chosen temperatures 500, 600, and

700 K. The diffusivities derived from mean square displacement are fitted to the Arrhenius

equation, and the results are shown in Fig. 1(b).

Fig. 1(a) indicates that the hydrogen atom mean square displacement satisfies well the

linear relationship with time. The statistical error of the t = 4.4 ps data is more significant than

the t = 0.0088 ps data. This is expected because the t = 4.4 ps case involves much less terms

(MD outputs) in the average sum. Considering the relatively large error bars, however, it is

somewhat surprising that the median points of the t = 4.4 ps data can fall almost right on the

linear lines. It should be noted that the error bars are defined here as the standard deviation of the

mean square displacement computed through Eq. (1). This definition is based on the assumption

that the error satisfies a normal distribution. It is unclear if this assumption is valid for Eq. (1).


According to Arrhenius equation D = D0 exp(-Q/kT), the normalized temperature T/Q reflects a material-
independent temperature effect.

7
Nonetheless, it is important to note that the good fits (of the median points) are highly

reproducible in our simulations (listed in Table I below) with change of conditions so that they

are not due to statistical reasons.

Fig. 1. (a) Mean square displacement as a function of time and (b) the Arrhenius plot for two
sets of MD simulations conducted using the NVT ensemble at two different temperature
ranges and t values, but the same time step size of dt = 0.001 ps and total MD time of
tMD = 0.88 ns.

To quantify the quality of the fits shown in Fig. 1(b), we define an error parameter for the

Arrhenius fit:

n
1
 
n  lnD   lnD
i 1
i
2
(2)

where n is total number of MD diffusivities Di (i = 1, 2, …, n), and D is the diffusivity calculated

from the fitted Arrhenius equation. Based on this definition, we find a  = 0.5188 for the t =

0.0088 ps data, and a  = 0.0759 for the t = 4.4 ps data. Now we can see two interesting

phenomena from Fig. 1(b). First, while the t = 4.4 ps data is associated with larger statistical

errors than the t = 0.0088 ps data for the mean square displacement as shown in Fig. 1(a), the fit

to the Arrhenius equation is not worse for the t = 4.4 ps data than for the t = 0.0088 ps data (in

8
fact, better at t = 4.4 ps with  = 0.0759 than at t = 0.0088 ps with  = 0.5188). To understand

this, we count number of hydrogen jumps in one of our simulations. We find that the hydrogen

residence time (inverse of jump frequency) is about 1.4 ps at 800 K and > 200 ps at 400 K.

Because t = 4.4 ps is comparable to resident time, the mean square displacement calculations

involve more diffusion component especially at the high temperature end. In contrast, t =

0.0088 ps is much shorter than the residence time, and therefore the mean square displacement

calculations are dominated by atom vibration rather than diffusion. Hence, reducing time steps

significantly below the residence time does not improve the accuracy of diffusion calculations.

The second interesting phenomenon in Fig. 1(b) is that the two data points at very low

temperature (1/kT > 24) seem to deviate from the linear function more significantly than the

other points. This temperature effect can be understood because at low temperatures, the number

of atom jumps is low resulting in statistical error due to insufficient samples. High temperatures

may also pose a problem because an extremely high temperature is equivalent to a barrier-less

diffusion (i.e., jump frequency approaches attempt frequency). Then the simulation would be

sampling the ballistic motion of the hydrogen atom rather than sensitively reflecting the

thermally activated diffusion, especially considering the small hydrogen mass. Hence, it is

always safer to apply an intermediate temperature range for calculating the diffusion energy

barrier using MD. The temperature range we use for the t = 4.4 ps simulations, T = 500 - 700 K

(T/Q = 1220 – 1707 K/eV), appears to be a good choice from Fig. 1(b).

Despite different values of t, the statistical errors of both sets of data shown in Fig. 1(b)

are too large to yield confident energy barrier calculations. We now explore the convergence of

the activation energy as a function of total MD time tMD where the other parameters are held

fixed at the better values determined above: dt = 0.001 ps, t = 4.4 ps, and T = 500, 525, 550,

9
575, 600, 625, 650, 675, and 700 K (T/Q = 1220 – 1707 K/eV). The Arrhenius error , and the

corresponding activation energy Q, are calculated at different tMD, and the results are shown in

Figs. 2(a) and 2(b) respectively. For comparison, the Arrhenius errors obtained from the data

shown in Fig. 1(b), and the activation energy obtained from MS simulations, are indicated in

Figs. 2(a) and 2(b) respectively using dash lines.

Fig. 2. (a) Arrhenius error  and (b) activation energy Q as a function of MD time tMD, obtained
from MD simulations using the NVT ensemble at a time step size of dt = 0.001 ps, an
output time interval of t = 4.4 ps, and nine temperatures T = 500, 525, 550, 575, 600,
625, 650, 675, and 700 K (T/Q = 1220 – 1707 K/eV).

Fig. 2(a) more clearly indicates that at tMD = 0.88 ns, increasing t from 0.0088 to 4.4 ps

greatly reduces the error of the Arrhenius fit. At the given t = 4.4 ps, increasing tMD beyond

0.88 ns results in a systematic decrease of the Arrhenius error. In particular, the Arrhenius error

becomes near zero (on the scale of the figure) when tMD reaches about 7 ns and above.

Correspondingly, Fig. 2(b) indicates that the activation energy is extremely sensitive to tMD when

tMD is small, but becomes insensitive to tMD at large tMD. In particular, the activation energy

nearly saturates when tMD reaches 7 ns. Interestingly, the saturated activation energy shown in

Fig. 2(b) is very close to the value obtained from MS simulations. Fundamentally, the MD data

10
should be taken as the standard if the statistical error is not large. This is because MD

incorporates the temperature effect (also the mass effect as will be discussed below), and reflects

the overall behaviour of all possible diffusion paths (e.g., O-T and T-O two-way jumps).

The discussions above indicate that a very low Arrhenius error can be achieved from MD

simulations using the NVT ensemble at dt = 0.001 ps, t = 4.4 ps, tMD = 13 ns, and nine

temperatures T = 500, 525, 550, 575, 600, 625, 650, 675, and 700 K. We now directly visualize

the mean square displacement and the Arrhenius fits for such simulations in Fig. 3. Fig. 3(a)

confirms that the mean square displacement has very small statistical errors and its median points

satisfy very well the linear relationship with time for all the nine temperatures. More importantly,

Fig. 3(b) indicates that the MD diffusivities satisfactorily fit the Arrhenius equation. This is an

important result indicating that indeed MD can be a robust method to study diffusion, which is

not limited by the complexity of the system.

Fig. 3. (a) Mean square displacement as a function of time and (b) the Arrhenius plot for MD
simulations conducted using the NVT ensemble at a time step size of dt = 0.001 ps, an
output time interval of t = 4.4 ps, a total MD time of tMD = 13.2 ns, and nine
temperatures T = 500, 525, 550, 575, 600, 625, 650, 675, and 700 K.

11
C. Parameter Study

Having established MD as a viable method for diffusion calculations, questions arise as

how results depend on the many MD simulation parameters such as ensembles, thermostat

methods, time step size, damping and dragging parameters for the thermostat and barostat, etc.

To understand these, extensive MD simulations are performed to calculate diffusion activation

energies at different ensembles and MD parameters. According to discussion above, all of our

calculations use appropriately large values of t ( 2.2 ps) and tMD ( 6.6 ns), and nine

temperatures T = 500, 525, 550, 575, 600, 625, 650, 675, and 700 K.

By solving atom positions entirely from Newton’s equations, the NVE ensemble does not

introduce additional parameters other than dt, t, tMD and T. The NVT ensemble, based on either

the Nose-Hoover [11] or the Langevin [12] thermostats, requires additional thermostat

parameters [9]. For the Nose-Hoover thermostat, these include the temperature damping

parameter (Tdamp), number of sub-cycles to perform thermostat within each time step (Tloop), and

the drag factor (drag). For the Langevin thermostat, the parameters include Tdamp, the switch

parameter for the Gronbech-Jensen/Farago (gjf = yes or no) algorithm [40,41], and the switch

parameter to set the total random force to zero (zero = yes or no). The NPT ensemble is based on

the Nose-Hoover barostat thermostat. In addition to Tdamp, Tloop, and drag mentioned above, it

also involves pressure damping parameter (Pdamp), number of sub-cycles to perform barostat

within each time step (Ploop), the number of steps for resetting the reference cell (nreset), and the

switch parameter to use Martyna-Tuckerman- Klein correction [42,43] (mtk = yes or no). For

NVT or NPT ensembles, there are three additional conditions that can be specified: A, the

thermostat is only applied to aluminium (i.e., leaving the hydrogen to the NVE ensemble) and

the initial velocity of the hydrogen atom is set according to the simulated temperature T; B, the

12
thermostat is only applied to aluminium and the hydrogen initial velocity is set to zero; and C,

the thermostat is applied to both aluminium and hydrogen but the hydrogen initial velocity is set

to T. Finally, the mass of the diffusion species can also be treated as a parameter so that we can

explore the effects of mass (for example, effects of hydrogen isotopes). With the MD parameters

defined, we now point out that the results discussed in Figs. 1-3 are obtained from an NVT

ensemble with a Nose-Hoover thermostat applied only to aluminium (using condition B) at Tdamp

= 100dt, Tloop = 1, and drag = 0.0.

NVE does not have a thermostat or a barostat so that care is made to ensure correct NVE

simulations. First, due to the equal partition of kinetic and potential energies, the initial

temperature of the system is created at the value twice of the target temperature. After ignore the

warm-up MD simulation, the time averaged temperature during the entire tMD simulation period

is taken as the true temperature. The true temperatures slightly differ from the target

temperatures (e.g., 485 vs. 500 K). We can of course repeat the simulation with the initial

temperature scaled up correspondingly so that the true temperature matches exactly the target

temperature. This, however, is not necessary because the analyses use only the true temperatures.

Second, we do verify that the total energy is conserved extremely well (e.g., statistically changes

from -6618.68 to -6618.50 eV, which can be attributed to numerical noises).

18 series of MD simulations are performed to examine effects of each MD parameter on

activation energy barrier Q, pre-exponential factor D0, and the Arrhenius error . In each series,

one or several related MD parameters are varied whereas the other parameters are kept constant.

Some parameters are changed together. For instance, when the number of steps are the same,

increasing dt would cause a corresponding increase in t and tMD. Under the conditions Tdamp =

100dt and Pdamp = 1000dt (which are normally used in MD simulations [9]), increasing dt would

13
also cause a corresponding increase in Tdamp and Pdamp. The results of all of the 18 series of

simulations are summarized in Table I, where different ensembles are separated by double lines,

and different series are separated by single lines and are marked by S#. We emphasize that the

results in Table I are not trivial, involving nearly 2000 simulations. This is because each

activation energy value is derived from nine temperatures, and each temperature consists of three

simulations with different random number seeds to enable us to achieve the total tMD within 1/3

of the wall clock time.

Table I. Numerical values of pre-exponential factor D0, activation energy barrier Q, and error of
the Arrhenius fit for all MD parameters explored. Note that different ensembles are
separated by double lines, and different series are separated by single lines and are
marked by S#. Here a series explores effects of targeted MD parameters while keeping
other parameters constant.
NVE parameters (mH=1amu) D0 Q 
S# dt (ps) t (ps) tMD (ns) (Å2/ps) (eV)
1 0.00075 3.3 9.9 73 0.38 0.0044
0.00100 4.4 13.2 81 0.39 0.0124
0.00125 5.5 15.5 83 0.39 0.0093
NVT parameters (Nose-Hoover) D0 Q 
S# dt (ps) t tMD Tdamp Tloop drag mH other* (Å2/ps) (eV)
(ps) (ns) (ps) (amu)
2 0.00050 2.2 6.6 0.050 1 0.0 1 A 203 0.44 0.0110
0.00075 3.3 9.9 0.075 1 0.0 1 A 161 0.42 0.0145
0.00100 4.4 13.2 0.100 1 0.0 1 A 149 0.42 0.0121
0.00125 5.5 15.5 0.125 1 0.0 1 A 107 0.40 0.0095
3 0.00075 3.3 9.9 0.075 1 0.0 1 B 160 0.42 0.0096
0.00100 4.4 13.2 0.100 1 0.0 1 B 180 0.43 0.0040
0.00125 5.5 15.5 0.125 1 0.0 1 B 121 0.41 0.0133
4 0.00075 3.3 9.9 0.075 1 0.0 1 C 146 0.42 0.0085
0.00100 4.4 13.2 0.100 1 0.0 1 C 137 0.42 0.0064
0.00125 5.5 15.5 0.125 1 0.0 1 C 138 0.42 0.0166
5 0.00075 3.3 9.9 0.010 1 0.0 1 A 147 0.42 0.0151
0.00100 4.4 13.2 0.010 1 0.0 1 A 80 0.39 0.0138
0.00125 5.5 15.5 0.010 1 0.0 1 A 68 0.38 0.0023
6 0.00050 2.2 6.6 0.005 1 0.0 1 A 92 0.39 0.0145
0.00050 2.2 6.6 0.010 1 0.0 1 A 225 0.44 0.0107
0.00050 2.2 6.6 0.030 1 0.0 1 A 151 0.42 0.0215
0.00050 2.2 6.6 0.050 1 0.0 1 A 203 0.44 0.0110
0.00050 2.2 6.6 0.070 1 0.0 1 A 121 0.41 0.0061

14
0.00050 2.2 6.6 0.090 1 0.0 1 A 140 0.42 0.0082
0.00050 2.2 6.6 0.110 1 0.0 1 A 371 0.47 0.0187
7 0.00050 2.2 6.6 0.050 1 0.0 1 A 203 0.44 0.0110
0.00050 2.2 6.6 0.050 2 0.0 1 A 60 0.37 0.0296
0.00050 2.2 6.6 0.050 3 0.0 1 A 114 0.41 0.0108
0.00050 2.2 6.6 0.050 4 0.0 1 A 70 0.38 0.0043
0.00050 2.2 6.6 0.050 5 0.0 1 A 389 0.47 0.0239
8 0.00050 2.2 6.6 0.050 1 0.0 1 A 203 0.44 0.0110
0.00050 2.2 6.6 0.050 1 0.2 1 A 87 0.40 0.0330
0.00050 2.2 6.6 0.050 1 0.4 1 A 220 0.44 0.0258
0.00050 2.2 6.6 0.050 1 0.6 1 A 188 0.43 0.0073
0.00050 2.2 6.6 0.050 1 0.8 1 A 79 0.39 0.0134
0.00050 2.2 6.6 0.050 1 1.0 1 A 179 0.43 0.0202
0.00050 2.2 6.6 0.050 1 1.2 1 A 101 0.40 0.0211
0.00050 2.2 6.6 0.050 1 1.4 1 A 64 0.37 0.0133
9 0.00050 2.2 6.6 0.050 1 0.0 1 A 203 0.44 0.0110
0.00050 2.2 6.6 0.050 1 0.0 2 A 103 0.41 0.0031
0.00050 2.2 6.6 0.050 1 0.0 4 A 94 0.42 0.0083
0.00050 2.2 6.6 0.050 1 0.0 6 A 26 0.36 0.0064
0.00050 2.2 6.6 0.050 1 0.0 8 A 24 0.36 0.0044
0.00050 2.2 6.6 0.050 1 0.0 10 A 26 0.37 0.0046
NVT parameters (Langevin, other=A, mH=1amu) D0 Q 
S# dt (ps) t tMD Tdamp gjf zero (Å2/ps) (eV)
(ps) (ns) (ps)
10 0.00050 2.2 6.6 0.050 yes yes 75 0.38 0.0033
0.00075 3.3 9.9 0.075 yes yes 124 0.41 0.0029
0.00100 4.4 13.2 0.100 yes yes 116 0.41 0.0017
0.00125 5.5 15.5 0.125 yes yes 166 0.42 0.0030
11 0.00050 2.2 6.6 0.030 yes yes 84 0.39 0.0019
0.00050 2.2 6.6 0.040 yes yes 111 0.40 0.0053
0.00050 2.2 6.6 0.050 yes yes 75 0.38 0.0033
0.00050 2.2 6.6 0.060 yes yes 90 0.40 0.0077
0.00050 2.2 6.6 0.070 yes yes 82 0.39 0.0079
12 0.00050 2.2 6.6 0.050 yes yes 75 0.38 0.0033
0.00050 2.2 6.6 0.050 no yes 42 0.36 0.0040
13 0.00050 2.2 6.6 0.050 yes yes 75 0.38 0.0033
0.00050 2.2 6.6 0.050 yes no fail fail fail
NPT parameters (Tdamp=100dt, Tloop=1ps, drag=0, other=A, D0 Q 
mH=1amu) (Å2/ps) (eV)
S# dt (ps) t tMD Pdamp Ploop nreset mtk
(ps) (ns) (ps)
14 0.00050 2.2 6.6 0.50 1 0 yes 72 0.39 0.0129
0.00075 3.3 9.9 0.75 1 0 yes 98 0.40 0.0120
0.00100 4.4 13.2 1.00 1 0 yes 137 0.42 0.0057
0.00125 5.5 15.5 1.25 1 0 yes 103 0.40 0.0081

15
15 0.00050 2.2 6.6 0.30 1 0 yes 77 0.39 0.0448
0.00050 2.2 6.6 0.40 1 0 yes 44 0.36 0.0557
0.00050 2.2 6.6 0.50 1 0 yes 72 0.39 0.0129
0.00050 2.2 6.6 0.60 1 0 yes 192 0.44 0.0134
0.00050 2.2 6.6 0.70 1 0 yes 190 0.44 0.0172
16 0.00050 2.2 6.6 0.50 1 0 yes 72 0.39 0.0129
0.00050 2.2 6.6 0.50 3 0 yes 75 0.39 0.0131
0.00050 2.2 6.6 0.50 5 0 yes 79 0.39 0.0316
0.00050 2.2 6.6 0.50 7 0 yes 190 0.44 0.0250
0.00050 2.2 6.6 0.50 9 0 yes 181 0.43 0.0355
17 0.00050 2.2 6.6 0.50 1 0 yes 72 0.39 0.0129
0.00050 2.2 6.6 0.50 1 2200 yes 81 0.39 0.0242
0.00050 2.2 6.6 0.50 1 4400 yes 108 0.41 0.0105
0.00050 2.2 6.6 0.50 1 6600 yes 59 0.37 0.0163
0.00050 2.2 6.6 0.50 1 8800 yes 67 0.38 0.0214
0.00050 2.2 6.6 0.50 1 11000 yes 115 0.41 0.0101
18 0.00050 2.2 6.6 0.50 1 0 yes 72 0.39 0.0129
0.00050 2.2 6.6 0.50 1 0 no 105 0.40 0.0081
*
: A: thermostat applies only to aluminum, hydrogen initial velocity not zero; B: thermostat
applies only to aluminum, hydrogen initial velocity zero; C: thermostat applies to all atoms,
hydrogen initial velocity not zero.

Table I provides significant insights on MD simulations of diffusion. First, the error of the

Arrhenius fit is significantly reduced in all of the MD simulations presented in Table I as

compared to the  = 0.5188 and  = 0.0759 of the shorter time MD simulations shown in Fig. 1.

As  is reduced, the activation energy Q increasingly converges towards ~0.41 eV (except for the

mass effect). According to Fig. 2,  can always be reduced by increasing tMD. Hence, Table I

strongly indicates that MD is a powerful and accurate method to calculate activation energy

barrier of diffusion where converged results independent of the MD parameters can be achieved

with a sufficiently long tMD.

S#9 indicates that increasing the mass of the diffusion species causes a systematic decrease

in both the pre-exponential factor D0 and activation energy barrier Q. The effects of mass on D0

can be understood because D0 scales with vibration frequency and vibration frequency reduces

with increasing mass. On the other hand, within the mass range of the three hydrogen isotopes

16
(hydrogen = 1 amu, deuterium = 2 amu, and tritium = 3 amu), the diffusion barriers are not too

different. Hence, one strategy to calculate diffusion barriers for hydrogen is to artificially

increase its mass to 2 or 3 amu because unlike mH = 1 amu which produces an error of  =

0.0110, increasing the mass above 2 seem to consistently lead to errors below the 0.01 margin at

the same tMD. Note that MD simulations of the light hydrogen atoms must use extremely small

time step size dt (e.g., dt << 0.001 ps). The reason that this problem does not occur in our case is

because we are only dealing with one hydrogen atom. Hence, increasing the mass has an

additional advantage for enabling a larger dt (and therefore a larger tMD for the same

computational cost) to be used to simulate systems containing many hydrogen atoms.

The NVE simulations (S# 1) seem to give very consistent D0 and Q values at various dt.

Parameters drag (S# 8), Tdamp (S# 6, 11), Tloop (S# 7), Pdamp (S# 15), Ploop (S# 16), nreset (S# 17),

and mtk (S# 18) also do not change the Arrhenius error or the activation energy barrier in a

systematic way, although  is noticeably larger when Pdamp is considerably lower than 1000dt

(e.g., Pdamp = 0.3 or 0.4 ps in S# 15). The default values drag = 0, Tdamp = 100dt, Tloop = 1, Pdamp

1000dt, Ploop = 1, nreset = 0, and mtk = yes [9] are good, although researchers can always make

judicious choices based on Table I.

The Langevin thermostat of the NVT ensemble (S# 10 - 12) produces systematically lower

Arrhenius error than other approaches. Hence, Langevin thermostat, in conjunction with the

NVT ensemble, is the preferred method to study diffusion with minimum computational cost

(minimum tMD). When using the Langevin thermostat, the total random force must be set to zero,

or otherwise the simulation (S# 13) completely fails. This means that the D vs. T data is too

irregular and scattered to fit a positive D0. On the other hand, the gjf correction of the Langevin

thermostat is good because the simulation that does not use this correction (S# 12) produces

17
noticeably low D0 and Q values as compared with the data obtained from all other simulations.

The only problem is that the Langevin thermostat is used with the NVT ensemble, and hence it is

not applicable to processes where changes of dimension and phase transformation occur.

At default values of drag = 0, Tdamp = 100dt, Tloop = 1, Pdamp 1000dt, Ploop = 1, nreset = 0, and

mtk = yes, the NPT calculations consistently give small errors of  < 0.013 (S# 14-18). Note that

NPT is only needed to address structural changes of the system, which can only be caused by

more than one reacting (diffusion) atoms. In that case, the number of samples‡ obtained from a

single simulation is multiplied by the number of diffusion atoms, and therefore, there are no

problems to achieve a much more converged calculation of the diffusion properties. In fact, our

NPT simulations of hydrogen diffusion in a PdH0.4 hydride lead to an Arrhenius error of  <

0.0000.

The pre-exponential factor D0 is not as converged as Q. In addition, there are no systematic

trends of D0 with the MD parameters, suggesting that the origin of the variation comes from

statistics. This is further verified as the variation of D0 falls within a narrower range of ~ 70-190

Å2/ps if only data with  < 0.01 (except due to the mass effect in S# 9, or due to the neglect of

the gjf effect in S# 12) is considered. Considering that D0 is an extrapolated quantity with

significant variation when measured experimentally [19,23], the range between 70-190 Å2/ps is

not too bad. On the other hand, it can be expected from Fig. 2 that converged D0 results can be

achieved when tMD is significantly increased. This is again verified by MD simulations of various

palladium hydrides PdH0.1, PdH0.2, PdH0.3, PdH0.4, PdH0.5, PdH0.6, PdH0.7, where D0 falls

narrowly between 33 and 37 Å2/ps. Note that in this case, the higher convergence is achieved by


Here a sample refers to a measurement of position of a diffusion atom.

18
more hydrogen atoms, which is equivalent to increasing total tMD when only one atom is

sampled.

Finally, conditions A, B, and C (i.e., if the hydrogen atom is subject to the thermostat and if

it has an initial velocity) do not seem to significantly impact the results (see, for example, S# 2-

4).

IV. EXPERIMENTAL COMPARISON

Due to the low solubility, the diffusion energy barrier of hydrogen in aluminium is difficult

to measure, and hence the reported values vary from 0.17 [25] to 1.45 [26] eV. To compare

experiments on a fair footing, we summarize in Table II the experimental results available in

literature.

Table II. Literature experimental values of diffusion energy barrier Q and pre-exponential factor
D0 for hydrogen in aluminium.
feature explored feature metric D0 (Å2/ps) Q (eV) literature
grain > 25 mm 9.530102 0.46 Ichimura et al [13,14]
size 15 mm 4.580102 0.38
4 mm 1.520103 0.55
2 mm 2.110104 0.69
3 mm 1.540105 0.84
3 mm 4.220106 1.11
4 mm 6.000105 0.92
Al purity 99.999 % 1.900103 0.41 Papp et al [15,16]
99.999 % + oxides 1.100105 0.74
(water-treated)
99.8 % 2.510106 0.93
unexplored ----- 1.100103 0.42 Eichenauer et al [17,18]
----- 2.100103 0.47
unexplored ----- 1.101103 0.49 Outlaw et al [19]
unexplored ----- 2.000102 0.52 Matsuo et al [20]
unexplored ----- 9.200103 0.57 Ishikawa et al [21]
unexplored ----- 6.100103 0.57 Saitoh et al [22]
unexplored ----- 2.600103 0.61 Hashimoto et al [23]
unexplored ----- 1.300105 0.69 Csanady et al [24]
unexplored ----- 1.750100 0.17 Young et al [25]

19
unexplored ----- 1.200109 1.45 Ransley et al [26]

Table II confirms that the literature data for hydrogen diffusion in aluminium is extremely

scattered. However, Ichimura et al [13,14] studied the effects of grain boundaries and Papp et al

[15,16] studied the effects of aluminium purity. Their results indicate that for pristine materials

with increasing grain size and purity, the hydrogen diffusion energy barrier increasingly

approaches the 0.41 – 0.46 eV margin, in good agreement with our results.

Interestingly, Table II also clearly shows a “compensation” effect [13], that is, large

activation energy barriers are always associated with large pre-exponential factors so that the

overall diffusion coefficients do not vary much. This suggests that the measured large energy

barriers pertain to traps rather than lattice sites. This is because to have a large overall diffusivity

when the barrier is high, the activated hydrogen atom must diffuse a long distance before being

de-activated again. Consistent with the this notion, the increase in energy barrier due to grain

boundaries and impurities has been interpreted by the trapping of hydrogen atoms to these

structural defects [13,25,27,29]. In this regard, it is the low end of the measured activation

energies that pertain to the lattice barrier calculated in our simulations. The lowest energy barrier

is 0.17 eV measured by Young et al [25]. This data point is “abnormally” lower than all the other

measurements. If this abnormal point is excluded, the lower end of the measured energy barriers

are again in the 0.38 – 0.46 eV range, in good agreement with our simulations. Confidence in the

simulations may also stem from the interatomic potential that captures a variety of Al-H

structures, as mentioned above.

We point out that our current studies cannot confidently address why the diffusion barrier

reported by Young et al [25] is significantly lower than our results and other experimental

measurements. On the other hand, the lowest energy barrier seen in experiments does not have to

20
be the lattice diffusion barrier. For instance, while grain boundaries can trap diffusion atoms

from the bulk, the trapped atoms can still diffuse along grain boundaries where the diffusion

barrier is in fact lower than in the bulk. Grain boundaries only increase diffusion barriers when

there are diffusion-stopping traps on the grain boundaries, such as the junction points of the grain

network [13]. By performing the same MD simulations on systems containing the relevant

defects, our methods will allow us to confidently study the effects of vacancies, grain

boundaries, and dislocations etc. in the future so that this issue can be directly addressed.

It is important to note that the experimental pre-exponential factors shown in Table II are in

general larger than the calculated values shown in Table I. Unlike diffusion energy barrier that

can be directly tied to the energetics of the interatomic potential, the discrepancy in pre-

exponential factor is less clear. Because experimental D0 is larger, there might be ballistic motion

of hydrogen atoms (i.e., jump length is longer than the lattice size) in experiments that is not

captured by simulations. Our highly converged results should establish an interesting topic for

future studies to clarify this issue.

V. CONCLUSIONS

We have demonstrated that robust MD methodologies can produce reliable diffusion

energy barriers from the Arrhenius equation and temperature-dependent diffusivity values. Most

importantly, such methods can be generally applied to complex materials including defective

metal alloys, hydrides, and oxides. Simulations using these methods lead to the following

insights:

1. The time interval t for tracking atom positions can be comparable or above the residence

time of diffusion species without impacting the convergence of the results, and highly

converged results can be achieved by extending the total simulated time tMD;

21
2. Parameters drag, Tdamp, Tloop, Pdamp, Ploop, nreset, and mtk do not impact results in a significant

way, whereas gjf = yes is desirable and zero = yes is mandatory for the Langevin thermostat.

Normally the default values of drag = 0, Tdamp = 100dt, Tloop = 1, Pdamp 1000dt, Ploop = 1, nreset

= 0, mtk = yes, gjf = yes, and zero = yes can be used, although researchers can always make

judicious choices based on Table I;

3. Increasing the mass of the diffusion atoms reduces both pre-exponential factor and diffusion

energy barrier, although the change of diffusion barrier is not significant for the mass range

of the three hydrogen isotopes (hydrogen, deuterium, and tritium);

4. NVE ensemble produces highly consistent results. When only simulating a single diffusion

atom, the NVT ensemble with a Langevin thermostat produces the lowest Arrhenius error as

compared with other ensembles. However, we find that the Arrhenius error of NPT ensemble

can be even lower if the system contains many diffusion atoms;

5. Our MD prediction of diffusion energy barrier of Q = 0.41 eV is in good agreement with the

0.38 – 0.46 eV experimental range for pristine samples with large grains and high purities.

The MD value also agrees with the MS calculations. Our study helps to elucidate the

inconsistence of the experimental results: the high measured energy barriers likely pertain to

traps, rather than lattice sites;

6. The use of a high fidelity Al-H bond order potential may contribute to the agreement

between our simulations and experiments.

VI. ACKNOWLEDGEMENTS

Sandia National Laboratories is a multiprogram laboratory managed and operated by Sandia

Corporation, a wholly owned subsidiary of Lockheed Martin Corporation, for the US

22
Department of Energy's National Nuclear Security Administration under contract DE-AC04-

94AL85000. This work was sponsored by DOE under HyMARC project.

References

1 L. Schlapbach, and A. Züttel, Nature, 414, 353 (2001).


2 R. Bhattacharyya, and S. Mohan, Ren. Sus. Ener. Rev., 41, 872 (2015).
3 K. Crosby, X. Wan, and L. L. Shaw, J. Power Sources, 195, 7380 (2010).
4 X. Chen, C. Li, M. Grätzel, R. Kostecki, and S. S. Mao, Chem. Soc. Rev., 41, 7909 (2012).
5 J. Zheng, X. Liu, P. Xu, P. Liu, Y. Zhao, and J. Yang, Inter. J. Hydro. Ener., 37, 1048
(2012).
6 G. Henkelman, and H. Jonsson, J. Chem. Phys., 113, 9978 (2000).
7 G. Henkelman, B. P. Uberuaga, and H. Jonsson, J. Chem. Phys., 113, 9901 (2000).
8 A. Nakano, Comp. Phys. Comm., 178, 280 (2008).
9 LAMMPS download site: lammps.sandia.gov.
10 S. Plimpton, J. Comp. Phys. 117, 1 (1995).
11 W. G. Hoover, Phys. Rev. B, 31, 1695 (1995).
12 T. Schneider, and E. Stoll, Phys. Rev. B, 17, 1302 (1978).
13 M. Ichimura, Y. Sasajima, and M. Imabayashi, Mater. Trans., JIM, 32, 1109 (1991).
14 M. Ichimura, and M. Imabayashi, Japan Inst. Metals, 43, 876 (1979).
15 K. Papp, and E. Kovacs-Csetenyi, Scr. Metall., 15, 161 (1981).
16 K. Papp, and E. Kovacs-Csetenyi, Scr. Metall., 11, 921 (1977).
17 W. Eichenauer, K. Hattenbach, and A. Pebler, Z. Metallk., 52, 682 (1961).
18 W. Eichenauer, and A. Pebler, Z. Metallk., 48, 373 (1957).
19 R. A. Outlaw, D. T. Peterson, and F. A. Schmidt, Scr. Metall., 16, 287 (1982).
20 S. Matsuo, and T. Hirata, J. Japan Inst. Metals, 31, 590 (1967).
21 T. Ishikawa, and R. B. McLellan, Acta Metall., 34, 1091 (1986).
22 H. Saitoh, Y. Iijima, and H. Tanaka, Acta Metall., 42, 2493 (1994).
23 E. Hashimoto, and T. Kino, J. Phys. F: Met. Phys., 13, 1157 (1983).
24 A. Csanady, K. Rapp, and E. Pasztor, Mater. Sci. Eng., 48, 35 (1981).
25 G. A. Young, and J. R. Scully, Acta Mater., 46, 6337 (1998).
26 C. E. Ransley, D. E. J. Talbot, Z. Metallk., 46, 328 (1955).
27 R. B. McLellan, Scr. Metall., 17, 1237 (1983).
28 S. Matuso, and T. Hirata, Trans. Nat. Res. Met., 11, 88 (1969).
29 R. A. H. Edwards, and W. Eichenauer, Scr. Metall., 14, 971 (1980).
30 H. Gunaydin, S. V. Barabash, K. N. Houk, and V. Ozolins, Phys. Rev. Lett., 101, 075901
(2008).
31 C. Wolverton, V. Ozoliņš, and M. Asta, Phys. Rev. B, 69, 144109 (2004).
32 J. Hirsch, and T. Al-Samman, Acta Mater., 61, 818 (2013).
33 T. Marlaud, B. Malki, C. Henon, A. Deschamps, and B. Baroux, Corr. Sci., 53, 3139 (2011).
34 H. Kamousti, G. N. Haidemenopoulos, V. Bontozoglou, and S. Pantelakis, Corros. Sci., 48,
1209 (2006).
35 T. Marlaud, B. Malki, A. Deschamps, and B. Baroux, Corros. Sci., 53, 1394 (2011).

23
36 T. Michler, C. S. Marchi, J. Naumann, S. Weber, and M. Martin, Inter. J. Hydro. Ener., 37,
16231 (2012).
37 X. W. Zhou, D. K. Ward, and M. E. Foster, submitted.
38 X. W. Zhou, D. K. Ward, M. Foster, J. A. Zimmerman, J. Mater. Sci., 50, 2859 (2015).
39 F. Reif, Fundamentals of Statistical and Thermal Physics (McGraw Hill, New York, 1965).
40 N. Gronbech-Jensen, and O. Farago, Mol. Phys., 111, 983 (2013).
41 N. Gronbech-Jensen, N. R. Hayre, and O. Farago, Comp. Phys. Comm., 185, 524 (2014).
42 G. J. Martyna, M. L. Klein, and M. Tuckerman, J. Chem. Phys., 97, 2635 (1992).
43 G. J. Martyna, M. E. Tuckerman, D. J. Tobias, and M. L. Klein, Mol. Phys., 87, 1117
(1996).

24

You might also like