Journal of Nuclear Materials 502 (2018) 161e168
Contents lists available at ScienceDirect
Journal of Nuclear Materials
journal homepage: www.elsevier.com/locate/jnucmat
Hydrogen retention in lithium and lithium oxide films
rrez b, A.O. Nelson c, M. Hofman d, P.S. Krsti
L. Buzi a, *, Y. Yang a, F.J. Domínguez-Gutie
c b,
c
a
R. Kaita , B.E. Koel
a
Department of Chemical and Biological Engineering, Princeton University, Princeton, NJ 08544-5263, United States
Institute for Advanced Computational Science, Stony Brook University, NY 11794-5250, United States
c
Princeton Plasma Physics Laboratory, Princeton, NJ 08543-0451, United States
d
Department of Chemistry, Princeton University, Princeton, NJ 08544, United States
b
a r t i c l e i n f o
a b s t r a c t
Article history:
Received 2 October 2017
Received in revised form
2 February 2018
Accepted 6 February 2018
Available online 9 February 2018
Pure lithium (Li) surfaces are difficult to maintain in fusion devices due to rapid oxide formation,
therefore, parameterizing and understanding the mechanisms of hydrogen (H, D) retention in lithium
oxide (Li2O) in addition to pure Li is crucial for Li plasma-facing material applications. To compare H
retention in Li and Li2O films, measurements were made as a function of surface temperature (90e520 K)
under ultrahigh vacuum (UHV) conditions using temperature programmed desorption (TPD). In both
cases, the total retention dropped with surface temperature, from 95% at 90 K to 35% at 520 K Li2O films
retained H in similar amounts as pure Li. Molecular Dynamics (MD) modeling was used to elucidate the
mechanisms of H retention, and results were consistent with experiments in terms of both retention
fraction and the drop of retention with temperature.
© 2018 Elsevier B.V. All rights reserved.
Keywords:
Hydrogen retention
Lithium
Lithium oxide
Nickel single crystal
Molecular dynamics
1. Introduction
Li conditioning of plasma facing components (PFCs) has
improved plasma performance with energy confinement and
lowered H recycling in magnetic fusion devices [1e4], and suppressed edge-localized modes (ELMs) in NSTX [5,6]. Li conditioning
of the NSTX divertor resulted in significant reduction (50%) of the
heat load due to enhanced radiation [7,8]. These effects may be due
in part to lithium's efficiency in binding H isotopes, thereby
increasing the H retention and lowering the recycling of these
species.
Accordingly, an understanding of H retention mechanisms and
parameterization of H uptake in Li is needed for future applications
of Li in high heat-flux and long-pulse duration machines, as well as
for H storage applications [9e11]. Such information is also needed
for Li2O because at typical base pressures of 1 10 8 Torr, which is
not uncommon in tokamaks, Li2O can form rapidly
(2Li þ H2O/Li2O þ H2 [12]). For instance, in NSTX the walls were
exposed to 100e600 L (1 L ¼ 1 10 6 Torr sec) of water vapor inbetween the plasma shots, during which the Li oxidation occurred.
* Corresponding author.
E-mail address:
[email protected] (L. Buzi).
https://doi.org/10.1016/j.jnucmat.2018.02.010
0022-3115/© 2018 Elsevier B.V. All rights reserved.
According to the residual gas analysis, the gas consisted of 77% of
hydrogenic species (mass 2, 3, 4) and 18% of water vapor (mass 17,
18, 19, 20) [13]. Li oxidation has also been observed in in-vacuo
measurements of Li-coated samples of PFCs in LTX using the Materials Analysis Particle Probe (MAPP) [14e16]. The problem will
continue to be important for the high-Z PFC phase of operation in
NSTX-U in which Li2O is also quickly formed, during Li evaporation,
under typical water partial pressure conditions in the 1 10 9 Torr
range.
Several studies have addressed the mechanisms of H retention
in Li. Baldwin et al., [17] measured deuterium (D) retention in Li as a
function of ion fluence and reported full uptake of D until volumetric conversion to LiD. Ion fluences beyond saturation led to a
switch from the low to the high recycling regime (i.e. high retention
to low retention), independent of the Li temperature (523e673 K).
Taylor et al., [18,19], after analyzing NSTX tiles and performing exsitu experiments on ATJ graphite samples, reported D bonding with
Li after Li interacted with carbon and oxygen. Krsti
c et al. demonstrated that O concentrations could increase up to 40 at% with
significant D fluence, which was later reconfirmed by experiments
by Taylor et al. [20]. Using quantum classical molecular dynamics
calculations (QCMD) on lithiated graphite surfaces, Krsti
c et al.
[21,22] showed that D is bound to O containing complexes rather
162
L. Buzi et al. / Journal of Nuclear Materials 502 (2018) 161e168
than to Li, thus promoting oxygen as having a main role in D
retention. MAPP results on LTX indicated that it is not crucial to
have just elemental Li to bind H, and Li oxide could also act as a
binding agent [16]. However, in order to gain a better understanding and evaluate Li oxide's ability to reduce recycling, the
efficiency of Li oxide compared to elemental Li in retaining H needs
further investigation.
This work is built upon previous experiments where an ultrathin (3 monolayer, ML) Li film was deposited on a polycrystalline
TZM sample and irradiated with D ions [23]. Release of oxygen from
TZM and subsequent oxidation of Li films during TPD, as well as the
complex Li-TZM interface, introduced uncertainties into interpreting those experiments. In order to address these issues, we
have conducted a systematic study in which five to seven times
thicker Li films were deposited on a nickel (Ni) single crystal. The
purity of the Li film was checked with auger electron spectroscopy
(AES). Modeling of the results is done by Molecular Dynamics (MD)
using REAXFF bond-order potentials [24,25], with correction for
dynamical polarization effects due to difference in Li, H and O
electronegativities by Electronegativity Equalization Method (EEM)
[26,27].
2. Experimental setup
All experiments were performed in a stainless steel ultrahigh
vacuum (UHV) chamber with a 2 10 10 Torr base pressure. Low
energy electron diffraction (LEED) was performed by using PHI
15e120 LEED optics. AES was done using a PHI 15-255G doublepass cylindrical mirror analyzer (CMA). TPD experiments were
performed with the sample in line-of-sight of the ionizer of a
shielded UTI 100C quadrupole mass spectrometer (QMS), with the
shield nozzle located 1 mm from the sample, using a heating rate of
4 K/s. A K-type thermocouple (chromel-alumel) was spot-welded
directly on the sample to monitor the temperature.
The Ni(110) single crystal (8 mm square, ±0.5 orientation)
sample was spot-welded onto tantalum (Ta) wires used for resistive
heating. The crystal was cleaned using cycles of 1.5-keV Arþ ion
sputtering and annealing in vacuum at 1100 K. Oxidation in
4 10 8 Torr O2 with the sample at 1000 K was used to eliminate
residual carbon, and reduction in 4 10 8 Torr H2 with the sample
at 1000 K was used to eliminate residual oxygen. Good surface order was confirmed with LEED. Surface purity was checked with AES
to ensure carbon and oxygen concentrations were below 1%. The
quality of the Ni(110) surface was also confirmed to be good using
the position and shape of the H2 desorption peaks in TPD [28]. In
these experiments, we used a Ni (110) single crystal as a substrate
to avoid effects due to grain boundaries, intrinsic defects, and impurities diffusing to the surface. Moreover, due to the low solubility
of alkali metals in Ni, Li and Ni are immiscible, and thus do not form
either bulk alloys or two-dimensional surface alloys [9,29].
Li dosing was performed with a commercial Li metal dispenser
(Li/NF/7.3/17/FT, SAES Group) by thermal evaporation onto the Ni
substrate. Hþ
2 ions were produced in a differentially pumped ion
gun (PHI 04e303 A) with adjustable ion energy from 0 to 5 keV, and
a liquid nitrogen trap was used in the H2 gas line to mitigate H2O
contamination. H2 gas (Praxair, 99.999%) and O2 gas (Praxair,
99.995%) was introduced into the chamber following a liquid nitrogen trap and using a high precision variable leak valve.
Hydrogen, rather than deuterium, was used in these experiments
for convenience and H has the same chemistry as D. After surface
preparation, Li films were exposed to a 500 eV Hþ
2 ion beam, which
þ
þ
is nominally composed of 90% Hþ
2 and 10% H [30]. The H2 flux was
defocused over the surface and the current measured on the
sample was 1.74 mA. The H2 pressure in the chamber during Hþ
2
irradiation was 4 10 8 Torr and the total exposure time was 120 s.
These conditions provided a total fluence of 4 1015 Hþcm
areal density was ~3 1015 Li cm 2, as discussed later.
2
. The Li
3. Computational approach
We simulated these experiments by using Molecular Dynamics.
Amorphous target surfaces of pure Li and Li2O were prepared for a
set of temperature values T (90, 300, 400, 500, and 600 K),
following the procedure in Refs. [21,22] for each temperature.
Computational cells of about 2000 atoms were used. These amorphous cells were created initially at 300 K, one with random distribution of lithium atoms, and another one with a predefined
random distribution of 33% O and 67% of Li atoms. These cells were
energy optimized in a succession of heating (1000 K max) and
annealing processes, and finally thermalized to a desired temperature using a Langevin thermostat with time constant of 100 fs. The
final numbers of atoms in the prepared cells at various temperatures are shown in Table 1.
Periodic boundary conditions were applied in the x-y directions
to simulate an infinite surface slab, with D impact in the z-direction.
The lateral dimensions of the cells were 3.6 nm in z direction and
about 3.4 nm in x and y directions for both surfaces (Fig. 1). The cell
depth of 3.6 nm is sufficient to prevent penetration of the D atoms
to the cell bottom boundary, thus avoiding artificial reflections.
The atomistic simulations were performed by MD, using Large
Scale Atomic/Molecular Massively Parallel Simulator (LAMMPS)
[31] with Reactive Force Field (ReaxFF) Bond Order (BO) potential
[24,25,32], and corrections for dynamic atom charge effects by
semi-empirical EEM [26,27]. The classical ReaxFF potentials used
for Li, O, D were verified in our previous computations [33] of
retention and sputtering of Li-C-O-D surfaces by Quantum-Classical
Molecular Dynamics (QCMD), using Self Consistent Charge Tight
Binding Density Functional Theory (SCC-DFTB) [34]. The ReaxFF
potentials implemented in LAMMPS are able to model the dynamics of breaking and forming of chemical bonds [24,32], as well
as to calculate the dynamic changes in charges of the atoms in the
system with the change of atomic coordinates using the Electronegativity Equalization Method [26,27]. The latter is important in
the presence of mutually polarizable materials with very different
electronegativities, such as Li (0.94) and O (3.4), while the H electronegativity (2.2) is in the middle.
The prepared computational cells, with various atomic contents
and at various temperatures, were bombarded by N ¼ 5040 independent 10 eV D atoms, with trajectories starting 1 nm at random
location above the surface, in the direction orthogonal to the surface. This large number of trajectories led to the adequate statistics
of H retention probability and was done at supercomputing facilities using “embarrassing parallelization”. If the number of D atoms
retained in the surface is ND, then the retention probability per D is
calculated as ND/N. The retention chemistry of D evolves at the end
of the collision cascade when the impact particle is thermalized
allowing comparison with the experimental results at higher
impact energies [21]. We carry out the analysis of the resulting
chemistry after the final rest location of each D impact, by performing the nearest-neighbor (NN) estimation for each retained D
(H) atom in the surface, defining the most-probable bonds [21].
While role of Li in bonding hydrogen is not challenged in the purelithium surface, it is surprising that O-D and Li-D NNs are similarly
represented. Although there is two times more Li than oxygen
atoms in the surface, oxygen has coordination number 2, two times
larger than coordination number of Li (1). This indicates approximate similar efficiency of H (D) retention in Li and Li2O surfaces. We
note that the O-D bond percentage is slightly bigger than Li-D one,
and that effectiveness of O to bond D slightly increases with
temperature.
163
L. Buzi et al. / Journal of Nuclear Materials 502 (2018) 161e168
Table 1
Atomic content for Li and Li2O target surfaces at various temperatures after energy optimization and thermalization processes.
Temperature (K)
Pure Li
Li2O
Li atoms
Li atoms
O atoms
90
305
400
500
600
2000
1340
660
2000
1340
660
1998
1335
658
1995
1329
656
1991
1320
658
Fig. 1. a) Li system (2000 lithium atoms), b) Li2O system: 33% of O (660 atoms) and 67% of Li (1340 atoms). Green and red symbols represent Li and O, respectively. (For interpretation of the references to colour in this figure legend, the reader is referred to the Web version of this article.)
4. Results and discussion
Auger electron spectroscopy (AES) analysis was done on the
sample surface to determine the elemental composition. Electrons
with energy 3 keV were used to irradiate the sample and the
generated Auger electrons were analyzed with a double-pass cylindrical mirror analyzer (CMA). Fast AES scans were taken on the
pure Li film to avoid oxidation and/or other impurity deposition.
AES spectra of the Li metal films on Ni (110) showed a Li peak at
51 eV, which is characteristic of pure metallic Li, as shown in Fig. 2.
Carbon and oxygen impurity concentrations in the film, as detected
by AES and also carbon monoxide (CO) TPD, were less than 1%.
When the Li film was oxidized by exposure to 10 L O2 at fixed
exposure temperatures in the range 90e520 K, the Li AES transitions split into two negative-going peaks at 36 and 42 eV, indicating
the formation of Li2O [9,35,36]. No metallic Li signal in AES
remained. The estimated Li film thickness was 5 nm; slightly below
the AES probing depth (~10 nm) therefore the Ni peaks are still
visible.
Fig. 2. AES of (top) Li metal and (bottom) Li2O films on Ni(110).
A total fluence of 4 1015 Hþ cm 2 of 500 eV Hþ
2 ions was used
to irradiate the Li and Li2O films after deposition. TPD was performed within 2e5 min after ion irradiation, time needed to move
the manipulator and place the sample in front of the mass spectrometer. Li and H2 TPD profiles are shown in Fig. 3.
The initial Li coverage before irradiation was checked with TPD
at room temperature and its reproducibility was within 1e2%.
During the Li and Hþ
2 exposures, the Ni substrate was kept at a fixed
exposure temperature, ranging from 90 to 520 K. Li and H2
desorption curves are given in Fig. 3 for each exposure temperature
in order to illustrate this influence, however we will focus our
discussion below on the two highlighted curves at 90 and 520 K. As
seen in Fig. 3(a), the amount of Li desorbing in the multilayer peak
near 600 K starts to decrease when exposure temperatures above
450 K are used, since Li starts evaporating significantly at temperatures above 450 K [37]. The higher temperature Li desorption
(700e1000 K), as shown more clearly on an expanded y scale in
Fig. 3(a1), corresponds to desorption from the Li monolayer [37]. Hþ
2
irradiation caused the formation of LiH and a new Li desorption
peak near 650 K due to LiH decomposition [9]. This Li peak is
coincident with the H2 desorption peak, shown in Fig. 3(b),
demonstrating that Li and H2 evolution in TPD is rate-limited by LiH
decomposition in the films, as previously reported [9]. The H2 TPD
curves show that the amount of H retained in the Li film at 90 K is
higher than that at 520 K and only a slight decrease was observed
from 300 K to 520 K. The amount is quantitatively analyzed in the
next section. No LiH (8 amu) desorption was detected in these
experiments.
To understand the H retention measurements, we need a
determination of the absolute coverage (atoms/cm2) for both Li and
H in the films. The Li coverage was calibrated by determining the Li
TPD area from a Li film that formed a saturation monolayer
coverage (just prior to formation of a Li multilayer desorption
feature) and assigning the integrated area under this Li TPD curve,
after correction for the mass spectrometer sensitivity to the
translational energy of the desorbed species [38], to a Li coverage
corresponding to a hexagonal close packed structure of Li adatoms
with their metallic atomic radii (3.02 1015 atoms/cm2) [37]. The
surface coverage (q) is given in ML, where 1 ML corresponds to the
164
L. Buzi et al. / Journal of Nuclear Materials 502 (2018) 161e168
LiOH (m/z ¼ 24) desorption was detected in TPD. TPD curves for
these desorbed products are provided in Fig. 4(aec). No desorption
of LiO, H2O, or LiH was observed. Fig. 4(a) shows that Li is retained
at the Ni surface to higher temperatures after oxidation, due to the
increased thermal stability of Li2O. In this case, Li is only desorbed
when Li2O decomposes, which produces a Li TPD peak at 900 K. The
temperature was increased up to 1200 K during TPD due to the
limitations with the thermocouple operation therefore the highest
temperature peak was not completely captured.
For 300 K exposure case (Fig. 4(a)), 18% of the Li layer was not
oxidized since a Li metallic peak at 570 K was still present (2.2 ML)
and about 10% of the Li layer (1.2 ML) formed the LiH (Li peak at
Fig. 3. (a) Li and (b) H2 TPD curves after pure Li films at 90e520 K were irradiated with
500 eV Hþ
2 . Panel (a1) shows data from panel (a) on an expended y scale.
surface Ni atom density of Ni(110) of 1.14 1015 atoms/cm2. This
gives an absolute Li coverage, qLi, of 2.56 ML referenced to the
Ni(110) surface atom density. The H coverage was determined in a
similar method by determining the integrated area under the H2
TPD profiles. We assign a surface saturation of qH ¼ 1.5 ML
(1.71 1015 atoms/cm2) when produced from dissociative adsorption of H2 on Ni(110) at 300 K [28,39].
Similar experiments to those described above were done for
Li2O films. These films were oxidized by the same exposure of 10 L
O2 at fixed exposure temperatures in the range 90e520 K. After Hþ
2
irradiation of Li2O films with Hþ
2 ions, H2 (m/z ¼ 2), Li (m/z ¼ 7) and
Fig. 4. (a) Li and (b) H2 TPD curves after Li2O films at 90e520 K were irradiated with
500 eV Hþ
2.
L. Buzi et al. / Journal of Nuclear Materials 502 (2018) 161e168
630 K). The exposure to 400 K produced a Li peak (LiH decomposition temperature) at 630 K (Fig. 4(a)), indicating that about 27% of
the Li film formed LiH (4.7 ML). H2 TPD showed a low temperature
peak at 630 K only for 90 K and 300 K exposure temperature
(Fig. 4(b)). When the temperature increased beyond 300 K, only the
high temperature peak remained at 900 K. It is not understood why
the H2 peak is missing for the 400 K exposure, and presumably this
is related to another channel for the consumption of H2 during or
prior to TPD, but this requires further studies. LiOH (m/z ¼ 24), Li2O
(m/z ¼ 30) and O2 (m/z ¼ 32) were also monitored and desorbed at
900 K, although the calibration for these species is not provided in
this work (see LiOH decomposition in Ref. [40]).
It is important to point out that no H was measured from TPD
experiments after irradiation with 500 eV Hþ
2 ions on Ni at room
temperature, therefore the hydrogen retention in these experiments can be attributed solely to the Li/Li2O layer. Energetic Hþ
2 ions
(500 eV) get implanted in the first few nm of the Ni substrate, but
since Ni does not retain H at room temperature and above, H
immediately diffuses out of Ni at these temperatures and reacts
with the Li/Li2O layer. In this case Ni serves as a reservoir (virtual
source) of thermal H, which interacts with Li/Li2O and forms other
compounds such as LiH, LiOH etc. From this point of view, the 10 eV
impact energy used in the MD calculation (in which case the impact
particles do not reach the bottom of the slab) is applicable to the
experimental results. This is supported by the fact that the retention chemistry develops mainly when impact particles thermalize
[21], irrespectively of their impact energy.
AES spectra were obtained from all of these surfaces at fixed
exposure temperatures in the range 90e520 K after Hþ
2 irradiation
and these are given in Fig. 5. As a reference, AES spectra from the
clean Ni(110) surface and pure Li on Ni(110) at 90 K is provided in
Fig. 5(a) and (a1). Also, AES spectra from the oxidized Li film surface
at 90 K prior to Hþ
2 irradiation is provided in Fig. 5(b) and (b1).
The top panels of Fig. 5 show AES spectra after Hþ
2 irradiation of
pure Li for the Li (a) and O (a1) regions. In Fig. 5(a), at 90 K, Hþ
2
irradiation causes no shift in the metallic Li peak at 51 eV; 500 eV
Hþ
2 irradiation of the 5 nm Li film mostly results in H implanted in
165
Ni, and at 90 K this H is not mobile, and remains primarily in the Ni
substrate, and so the metallic Li signal in AES is maintained. The
penetration depth in Ni for 500 eV Hþ
2 ions, calculated by the SRIM
code [41] is 6 nm and it scales up to 10 nm for the Li/Ni system.
Li normally attenuates the Ni peak at 61 eV, except on the curve
labeled “a”, where Li exposure occurred at 300 K. In the 300 K case,
the Li layer may have been slightly thinner. Alternatively, the “as
dosed Li” may have not formed a uniform layer at 300 K, exposing
some Ni, whereas at higher temperatures Li diffuses to form a more
uniform layer covering the Ni surface. The O AES spectra shown in
Fig. 5(a1) show a small oxygen peak at 512 eV, which is still less than
2% in the film.
The bottom panels of Fig. 5 provide AES spectra taken after Hþ
2
irradiation of Li2O films for (b) Li and (b1) O AES regions. Exposure
of a 5 nm thick Li film to 10 L O2 oxidizes the Li to create a Li2O film.
As shown in Fig. 5(b), this causes the metallic Li AES peak at 51 eV to
be eliminated and two peaks near 36 and 42 eV, characteristic of
Li2O, to appear. However, also LiH has peaks in the range 36e42 eV
[9,35]. In order to understand whether the peak represents LiH or
Li2O (peak at 40þ-1 eV), we look at the shift in the O peak [35,42].
At temperatures up to 470 K, O peaks appear at 508e509 eV
(LiOH), and at higher temperatures it shifts to 512 eV (Li2O). In the
literature [36] it is shown that slow conversion of LiOH to Li2O and
H2O occurs on the order of minutes. In the present experiments,
TPD was performed 2e5 min after Hþ
2 irradiation, therefore, at
temperatures higher than 470 K, it is believed that LiOH decomposed to form Li2O and H2O according to the reaction: 2LiOH> Li2O þ H2O (g). H2O formation between Hþ
2 exposure and TPD
may be another cause of the H retention drop at elevated
temperatures.
In Fig. 6, the total amounts of Li and H2 are plotted for both cases,
i.e., pure and oxidized Li. The data points are connected with a
spline fit to guide the eye and the dashed and dotted lines in the
graph show the initial Li coverage in Li and Li2O experiments before
Hþ
2 ion irradiation. For the oxidized Li film, the relative amount of Li
desorbed from the irradiated Li2O film is lower than for the Li film
indicating that the Hþ
2 sputtering of pure Li is higher than
þ
Fig. 5. a) Li and a1) O AES signal after Hþ
2 exposure of pure Li and b) Li and b1) O AES signal after H2 exposure of Li2O.
166
L. Buzi et al. / Journal of Nuclear Materials 502 (2018) 161e168
Fig. 6. Concentrations of H and Li from TPD measurements after irradiation of Li films.
The dashed and dotted lines show initial Li coverage in Li and Li2O experiments.
sputtering of Li2O. We note no large effect of exposure temperatures on H retention exceeding the Li melting point. This was to be
expected since the H retaining compounds, LiH and LiOH, decompose at much higher temperatures than pure Li (see Figs. 3 and 4).
These measurements were carried out below the saturation H uptake level (Li:H ¼ 1) to more clearly identify changes in retention in
the films due to exposure temperature.
Results from both experiments and MD simulations are
compared in Fig. 7(a) and (b). In Fig. 7(a), the retention fraction of H
is plotted for pure and oxidized Li and compared with computed
MD results for D. The total incident H was calculated by the TPD
profile of H retained in Ni (110) at 90 K (8.2 ML/Ni). The retention
for Li and Li2O is rather similar. We have to keep in mind that
retention in the case of oxidized Li may be higher than shown due
to the release of LiOH, which is not calibrated and not accounted for
in this graph. Noise/scattering in the data is attributed to the statistical error due to the integration of the TPD profiles, and the 10%
error bar is associated with the inherent uncertainty of the
measuring instruments and the reproducibility of Hþ
2 dosing. At
Fig. 7. H retention fraction (H retention/H incident) in Li and Li2O as a function of
exposure temperature and comparison with D and H retention from MD calculations.
temperatures higher than 450 K (Fig. 6(b)), Li coverage starts to
drop rapidly due to evaporation, which may also contribute to a
larger scatter in the data for H retention.
For each computed data point, we report a standard error
defined as sS ¼ sN 1/2, where s is the standard deviation of our
sample of N cases. In summary, the MD calculations demonstrated
that the oxidation of the Li slab results in the formation of various Li
oxides and the retained D atoms are located in the interstitial positions of Li or bonded to oxygen. We also calculate the retention of
the same samples but with impact of H. The values obtained for the
retention are 89 ± 2.5% at 90 K, 72 ± 3% at 300 K and 62 ± 3% at
500 K.
A comparison between the computed values of H and D retention and experimental results for H retention is provided in
Fig. 7(b). Experiments and calculations consistently show retention
by both pure Li and Li2O, indicating that both pure and oxidized Li
can bind H with similar efficiency. The same drop of retention with
Texp is followed for both Li and Li2O. The exponential decay of
retention with temperature may originate from diffusion, since it is
a thermally activated process (i.e., higher Texp leads to higher H
desorption from the Li surface.) An interesting observation was the
comparison between H and D in terms of total retention in pure Li
(Fig. 7(b)). From the calculations, one can see a small drop in H
retention compared to D. Although H and D are chemically similar,
their different mass could lead to different diffusion rates in Li.
As a final comment, a similar chemistry is expected for tritium
and since tritium inventory is recognized as an important challenge
for a fusion reactor, possible solutions are being considered (see
Ref. [44]).
5. Conclusions
H retention in pure Li and Li2O films was calculated by molecular
dynamics and measured experimentally as a function of surface
temperature under UHV conditions. It was experimentally shown
that upon oxidation, Li thermal stability increased. It was also
shown that both pure Li and Li2O are able to retain H in almost
same amounts. In addition, it was shown that H and D retention
drops with surface temperature in the range 90e520 K from 95% to
35% due to outwards diffusion of H at high temperatures.
Results of the MD modeling with EEM corrections were qualitatively consistent with experimental results in terms of both
retention fraction and the drop of retention with temperature,
when using either H or D as impact particles. Similar trends and
agreements were seen between the experimental results for Li2O
and the MD results for a mixture of Li and oxygen with atomic
concentration ratio Li:O ¼ 2:1.
In TPD measurements, the Li and H desorption peaks were
observed at the same temperature (650 K) and this was interpreted
as, corresponding to the decomposition of LiH. In addition to TPD
analysis for Li2O films irradiated with H ions, AES measurements
were consistent with the formation of LiOH, which decomposed to
Li2O and H2O at exposure temperatures higher than 470 K.
The results from experiment and MD modeling do not preclude
the possibility that the hydrogen is retained through the formation
of large LinHn molecules. Earlier DFT studies, for example, predicted
the formation of a rock-salt structure for LiD under deuterium
bombardment of lithium films [43]. The Li and H could then bind to
form LiH at the surface, which would subsequently dissociate at the
same temperature as observed experimentally.
The details of the chemistry underlying H retention and Li and
Li2O films may thus be complex. Nevertheless, both experiment and
modeling support the possibility that low H recycling can be achieved if Li2O is formed under fusion reactor conditions, as both Li
and Li2O have comparable efficiency for trapping H.
L. Buzi et al. / Journal of Nuclear Materials 502 (2018) 161e168
Note that the digital data for the figures in this paper can be
found at http://arks.princeton.edu/ark:/88435/dsp01x920g025r.
[14]
Acknowledgements
[15]
This material is based upon work supported by the U.S.
Department of Energy, Office of Science/Fusion Energy Sciences
under Award DE-SC0012890 (BEK) and Award DE-SC0013752
(PSK). FJDG acknowledges support by a Postdoctoral Scholarship
Number 267898 of the National Council of Science and Technology,
Mexico.
[16]
[17]
References
[18]
[1] D.K. Mansfield, K.W. Hill, J.D. Strachan, M.G. Bell, S.D. Scott, R. Budny,
E.S. Marmar, J.A. Snipes, J.L. Terry, S. Batha, R.E. Bell, M. Bitter, C.E. Bush,
Z. Chang, D.S. Darrow, D. Ernst, E. Fredrickson, B. Grek, H.W. Herrmann,
A. Janos, D.L. Jassby, F.C. Jobes, D.W. Johnson, L.C. Johnson, F.M. Levinton,
D.R. Mikkelsen, D. Mueller, D.K. Owens, H. Park, A.T. Ramsey, A.L. Roquemore,
C.H. Skinner, T. Stevenson, B.C. Stratton, E. Synakowski, G. Taylor, A. von Halle,
S. von Goeler, K.L. Wong, S.J. Zweben, Enhancement of tokamak fusion test
reactor performance by lithium conditioning, Phys. Plasmas 3 (1996) 1892,
https://doi.org/10.1063/1.871984.
[2] M.G. Bell, H.W. Kugel, R. Kaita, L.E. Zakharov, H. Schneider, B.P. LeBlanc,
D. Mansfield, R.E. Bell, R. Maingi, S. Ding, S.M. Kaye, S.F. Paul, S.P. Gerhardt,
J.M. Canik, J.C. Hosea, G. Taylor, Plasma response to lithium-coated plasmafacing components in the National Spherical torus experiment, Plasma Phys.
Contr. Fusion 51 (2009) 124054, https://doi.org/10.1088/0741-3335/51/12/
124054.
[3] R. Majeski, R. Doerner, T. Gray, R. Kaita, R. Maingi, D. Mansfield, J. Spaleta,
V. Soukhanovskii, J. Timberlake, L. Zakharov, Enhanced energy confinement
and performance in a low-recycling tokamak, Phys. Rev. Lett. 97 (2006) 1e4,
https://doi.org/10.1103/PhysRevLett.97.075002.
[4] J.C. Schmitt, R.E. Bell, D.P. Boyle, B. Esposti, R. Kaita, T. Kozub, B.P. LeBlanc,
M. Lucia, R. Maingi, R. Majeski, E. Merino, S. Punjabi-Vinoth, G. Tchilingurian,
A. Capece, B. Koel, J. Roszell, T.M. Biewer, T.K. Gray, S. Kubota, P. Beiersdorfer,
K. Widmann, K. Tritz, High performance discharges in the Lithium Tokamak
eXperiment with liquid lithium walls, Phys. Plasmas 22 (2015), https://
doi.org/10.1063/1.4921153.
[5] R. Maingi, S.M. Kaye, C.H. Skinner, D.P. Boyle, J.M. Canik, M.G. Bell, R.E. Bell,
T.K. Gray, M.A. Jaworski, R. Kaita, H.W. Kugel, B.P. Leblanc, D.K. Mansfield,
T.H. Osborne, S.A. Sabbagh, V.A. Soukhanovskii, Continuous improvement of
H-mode discharge performance with progressively increasing lithium coatings in the National Spherical Torus Experiment, Phys. Rev. Lett. 107 (2011),
https://doi.org/10.1103/PhysRevLett.107.145004.
[6] R. Maingi, D.P. Boyle, J.M. Canik, S.M. Kaye, C.H. Skinner, J.P. Allain, M.G. Bell,
R.E. Bell, S.P. Gerhardt, T.K. Gray, M. a. Jaworski, R. Kaita, H.W. Kugel,
B.P. LeBlanc, J. Manickam, D.K. Mansfield, J.E. Menard, T.H. Osborne, R. Raman,
A.L. Roquemore, S. a. Sabbagh, P.B. Snyder, V. a. Soukhanovskii, The effect of
progressively increasing lithium coatings on plasma discharge characteristics,
transport, edge profiles and ELM stability in the National Spherical Torus
Experiment, Nucl. Fusion 52 (2012) 83001, https://doi.org/10.1088/00295515/52/8/083001.
[7] M. Ono, M.A. Jaworski, R. Kaita, H.W. Kugel, J.-W. Ahn, J.P. Allain, M.G. Bell,
R.E. Bell, D.J. Clayton, J.M. Canik, S. Ding, S. Gerhardt, T.K. Gray,
W. Guttenfelder, Y. Hirooka, J. Kallman, S. Kaye, D. Kumar, B.P. Leblanc,
R. Maingi, D.K. Mansfield, A. McLean, J. Menard, D. Mueller, R. Nygren, S. Paul,
M. Podesta, R. Raman, Y. Ren, S. Sabbagh, F. Scotti, C.H. Skinner,
V. Soukhanovskii, V. Surla, C.N. Taylor, J. Timberlake, L.E. Zakharov, Recent
progress in the NSTX/NSTX-U lithium programme and prospects for reactorrelevant liquid-lithium based divertor development, Nucl. Fusion 53 (2013),
https://doi.org/10.1088/0029-5515/53/11/113030.
[8] T.K. Gray, J.M. Canik, R. Maingi, a. G. McLean, J.-W. Ahn, M. a. Jaworkski,
R. Kaita, M. Ono, S.F. Paul, The effects of increasing lithium deposition on the
power exhaust channel in NSTX, Nucl. Fusion 54 (2014) 23001, https://
doi.org/10.1088/0029-5515/54/2/023001.
[9] J. Engbæk, G. Nielsen, J.H. Nielsen, I. Chorkendorff, Growth and decomposition
of lithium and lithium hydride on nickel, Surf. Sci. 600 (2006) 1468e1474,
https://doi.org/10.1016/j.susc.2006.01.040.
[10] L. Wang, M.Z. Quadir, K.F. Aguey-Zinsou, Direct and reversible hydrogen
storage of lithium hydride (LiH) nanoconfined in high surface area graphite,
Int. J. Hydrogen Energy 41 (2016) 18088e18094, https://doi.org/10.1016/
j.ijhydene.2016.07.073.
[11] L. Wang, M.Z. Quadir, K.F. Aguey-Zinsou, Ni coated LiH nanoparticles for
reversible hydrogen storage, Int. J. Hydrogen Energy 41 (2016) 6376e6386,
https://doi.org/10.1016/j.ijhydene.2016.01.173.
[12] C. Haertling, R.J. Hanrahan, R. Smith, A literature review of reactions and kinetics of lithium hydride hydrolysis, J. Nucl. Mater. 349 (2006) 195e233,
https://doi.org/10.1016/j.jnucmat.2005.10.005.
[13] C.H. Skinner, R. Sullenberger, B.E. Koel, M.A. Jaworski, H.W. Kugel, Plasma
[19]
[20]
[21]
[22]
[23]
[24]
[25]
[26]
[27]
[28]
[29]
[30]
[31]
[32]
[33]
[34]
[35]
[36]
[37]
[38]
167
facing surface composition during NSTX Li experiments, J. Nucl. Mater. 438
(2013), https://doi.org/10.1016/j.jnucmat.2013.01.136.
C.N. Taylor, B. Heim, S. Gonderman, J.P. Allain, Z. Yang, R. Kaita,
A.L. Roquemore, C.H. Skinner, R.A. Ellis, Materials analysis and particle probe:
a compact diagnostic system for in situ analysis of plasma-facing components
(invited), Rev. Sci. Instrum. (2012), https://doi.org/10.1063/1.4729262.
M. Lucia, R. Kaita, R. Majeski, F. Bedoya, J.P. Allain, D.P. Boyle, J.C. Schmitt,
D.A.S. Onge, Development progress of the materials analysis and particle
probe, Rev. Sci. Instrum. 85 (2014), https://doi.org/10.1063/1.4890257.
R. Kaita, M. Lucia, J.P. Allain, F. Bedoya, R. Bell, D. Boyle, A. Capece, M. Jaworski,
B.E. Koel, R. Majeski, J. Roszell, J. Schmitt, F. Scotti, C.H. Skinner,
V. Soukhanovskii, Hydrogen retention in lithium on metallic walls from invacuo analysis in LTX and implications for high-Z plasma-facing components in NSTX-U, Fusion Eng. Des. 117 (2017) 135e139, https://doi.org/
10.1016/j.fusengdes.2016.06.056.
M.J. Baldwin, R.P. Doerner, S.C. Luckhardt, R.W. Conn, Deuterium retention in
liquid lithium, Nucl. Fusion 42 (2002) 1318e1323, https://doi.org/10.1088/
0029-5515/42/11/305.
C.N. Taylor, B. Heim, J.P. Allain, Chemical response of lithiated graphite with
deuterium irradiation, J. Appl. Phys. 109 (2011), https://doi.org/10.1063/
1.3555097.
C.N. Taylor, J.P. Allain, B. Heim, P.S. Krsti
c, C.H. Skinner, H.W. Kugel, Surface
chemistry and physics of deuterium retention in lithiated graphite, J. Nucl.
Mater. (2011), https://doi.org/10.1016/j.jnucmat.2010.09.049.
C.N. Taylor, J.P. Allain, K.E. Luitjohan, P.S. Krsti
c, J. Dadras, C.H. Skinner,
Differentiating the role of lithium and oxygen in retaining deuterium on
lithiated graphite plasma-facing components, Phys. Plasmas 21 (2014),
https://doi.org/10.1063/1.4874340.
P.S. Krsti
c, J.P. Allain, C.N. Taylor, J. Dadras, S. Maeda, K. Morokuma, J. Jakowski,
A. Allouche, C.H. Skinner, Deuterium uptake in magnetic-fusion devices with
lithium-conditioned carbon walls, Phys. Rev. Lett. 110 (2013), https://doi.org/
10.1103/PhysRevLett.110.105001.
P.S. Krsti
c, J.P. Allain, A. Allouche, J. Jakowski, J. Dadras, C.N. Taylor, Z. Yang,
K. Morokuma, S. Maeda, Dynamics of deuterium retention and sputtering of
Li-C-O surfaces, Fusion Eng. Des. 87 (2012) 1732e1736, https://doi.org/
10.1016/j.fusengdes.2011.07.009.
A.M. Capece, J.P. Roszell, C.H. Skinner, B.E. Koel, Effects of temperature and
surface contamination on D retention in ultrathin Li films on TZM, J. Nucl.
https://doi.org/10.1016/
Mater.
463
(2015)
1177e1180,
j.jnucmat.2014.10.048.
A.C.T. Van Duin, S. Dasgupta, F. Lorant, W.A. Goddard, ReaxFF: a reactive force
field for hydrocarbons, J. Phys. Chem. A 105 (2001) 9396e9409, https://
doi.org/10.1021/jp004368u.
D. Bedrov, G.D. Smith, A.C.T. Van Duin, Reactions of singly-reduced ethylene
carbonate in lithium battery electrolytes: a molecular dynamics simulation
study using the ReaxFF, J. Phys. Chem. A (2012) 2978e2985, https://doi.org/
10.1021/jp210345b.
W.J. Mortier, S.K. Ghosh, S. Shankar, Electronegativity-equalization method
for the calculation of atomic charges in molecules, J. Am. Chem. Soc. 108
(1986) 4315e4320, https://doi.org/10.1021/ja00275a013.
Y. Cong, Z.Z. Yang, General atom-bond electronegativity equalization method
and its application in prediction of charge distributions in polypeptide, Chem.
Phys. Lett. 316 (2000) 324e329, https://doi.org/10.1016/S0009-2614(99)
01289-0.
T.E. Jackman, K. Griffiths, W.N. Unertl, J.A. Davies, K.H. Gurtler,
D.A. Harrington, P.R. Norton, Surface phases of Ni(110) induced by adsorption
of deuterium, Surf. Sci. 179 (1987) 297e321, https://doi.org/10.1016/00396028(87)90060-4.
K. Wang, P. Ross, Reversible Li deposition on Ni in ultrahigh-vacuum,
J. Electrochem. Soc. 142 (1995) L95eL97, https://doi.org/10.1149/1.2044257.
H. Tawara, Y. Itikawa, H. Nishimura, M. Yoshino, Cross sections and related
data for electron collisions with hydrogen molecules and molecular ions,
J. Phys. Chem. Ref. Data 19 (1990) 617e636, https://doi.org/10.1063/1.555856.
S. Plimpton, Fast Parallel algorithms for short-range molecular dynamics,
J. Comput. Phys. 117 (1995) 1e19, https://doi.org/10.1006/jcph.1995.1039.
S.S. Han, A.C.T. van Duin, W.A. Goddard III, H.M. Lee, Optimization and
application of lithium parameters for the reactive force field, ReaxFF, J. Phys.
Chem. A 109 (2005) 4575e4582, https://doi.org/10.1021/jp051450m.
rrez, P.S. Krsti
F.J. Domínguez-Gutie
c, Sputtering of lithiated and oxidated
carbon surfaces by low-energy deuterium irradiation, J. Nucl. Mater. 492
(2017) 56e61, https://doi.org/10.1016/j.jnucmat.2017.05.014.
M. Elstner, D. Porezag, G. Jungnickel, J. Elsner, M. Haugk, T. Frauenheim,
S. Suhai, G. Seifert, Self-consistent-charge density-functional tight-binding
method for simulations of complex materials properties, Phys. Rev. B 58
(1998) 7260e7268, https://doi.org/10.1103/PhysRevB.58.7260.
R.E. Clausing, D.S. Easton, G.L. Powell, Auger spectra of lithium metal and
lithium oxide, Surf. Sci. 36 (1973) 377e379, https://doi.org/10.1016/00396028(73)90268-9.
G.L. Powell, O. Ridge, G.E. Mcguire, D.S. Easton, R.E. Clausing, M.C. Division,
O. Ridge, O. Ridge, N. Division, U.C. Corporation, L.A.S. Laboratories, L. Alamos,
Auger spectra of lithi~ hydride* 46 (1974) 345e357.
V. Saltas, C.A. Papageorgopoulos, Adsorption of Li on Ni(110) surfaces at low
and room temperature, Surf. Sci. 461 (2000) 219e230, https://doi.org/
10.1016/S0039-6028(00)00588-4.
Y.-L. Tsai, C. Xu, B.E. Koel, Chemisorption of ethylene, propylene and
168
L. Buzi et al. / Journal of Nuclear Materials 502 (2018) 161e168
isobutylene on ordered Sn/Pt(111) surface alloys, Surf. Sci. 385 (1997) 37e59.
http://ac.els-cdn.com/S0039602897001143/1-s2.0-S0039602897001143main.pdf?_tid¼f06a722e-7933-11e7-867a00000aab0f02&acdnat¼1501865144_fe246f0f393ad353fe1e76f3379fb1d6.
[39] B. Voigtl€
ander, S. Lehwald, H. Ibach, Hydrogen adsorption and the adsorbateinduced Ni(110) reconstruction- an EELS study, Surf. Sci. 208 (1989) 113e135,
https://doi.org/10.1016/0039-6028(89)90039-3.
[40] L.N. Dinh, C.M. Cecala, J.H. Leckey, M. Balooch, The effects of moisture on LiD
single crystals studied by temperature-programmed decomposition, J. Nucl.
Mater. 295 (2001) 193e204, https://doi.org/10.1016/S0022-3115(01)005487.
[41] J.F. Ziegler, M.D. Ziegler, J.P. Biersack, SRIM - the stopping and range of ions in
matter (2010), Nucl. Instrum. Methods Phys. Res. Sect. B Beam Interact. Mater.
Atoms 268 (2010) 1818e1823, https://doi.org/10.1016/j.nimb.2010.02.091.
[42] J.P. Tonks, M.O. King, E.C. Galloway, J.F. Watts, Corrosion studies of LiH thin
films, J. Nucl. Mater. 484 (2017) 228e235, https://doi.org/10.1016/
j.jnucmat.2016.12.008.
[43] M. Chen, T. Abrams, M.A. Jaworski, E.A. Carter, Rock-salt structure lithium
deuteride formation in liquid lithium with high-concentrations of deuterium :
a first-principles molecular dynamics study, Nucl. Fusion 56 (2016) 16020,
https://doi.org/10.1088/0029-5515/56/1/016020.
[44] C.A. Gentile, L.E. Zakharov, R.D. Woolley, C.J. Huber, Flowing Liquid Lithium for
the Purpose of Reducing Tritium Inventory Levels in Fusion Energy Reactors,
2013. http://ieeexplore.ieee.org/stamp/stamp.jsp?tp¼&arnumber¼6635397.