You are on page 1of 9

Letter

www.acsmaterialsletters.org

Developing Ideal Metalorganic Hydrides for


Hydrogen Storage: From Theoretical
Prediction to Rational Fabrication
Zijun Jing, Qinqin Yuan, Yang Yu, Xiangtao Kong, Khai Chen Tan, Jintao Wang, Qijun Pei,
Xue-Bin Wang,* Wei Zhou, Hui Wu,* Anan Wu,* Teng He,* and Ping Chen
Downloaded via COMSATS INST OF INFORMATION TECHLGY on September 2, 2021 at 06:05:29 (UTC).

Cite This: ACS Materials Lett. 2021, 3, 1417−1425 Read Online


See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ACCESS Metrics & More Article Recommendations *


sı Supporting Information

ABSTRACT: Materials for hydrogen storage have been


extensively explored for a few decades. Thousands of materials
have been synthesized and tested; however, few systems could
meet the practical requirements. Metalorganic hydrides
discovered recently offer new opportunities. It is, however,
extremely time-consuming and inefficient to experimentally
screen potential materials from a large variety of metal cations
and organic anions. In the present study, we performed wide-
ranging theoretical predictions and screened more than 90
metalorganic hydrides; 20 of them were identified with both
high hydrogen capacities (≥5 wt %) and suitable thermody-
namics (heat of H2 desorption: 25−35 kJ/mol-H2) that allow
hydrogen uptake and release near ambient condition. We
picked up four of them, i.e., Li/Na indolides and 7-azaindolides, for experimental validation. All the four candidates can be
easily synthesized via the reactions between the metal hydrides and the corresponding organic precursors. Among them the
structures of lithium indolide (P21/c (no. 14)) and sodium 7-azaindolide (P4̅21c (no. 114) were successfully resolved.
Photoelectron spectroscopy and quantum chemical calculations confirmed that the charge density of the organic ring
increased with the introduction of an alkali metal, thus optimizing their ΔHd for hydrogen storage. Importantly, we
demonstrated experimentally that lithium indolide with a theoretical hydrogen capacity of 6.1 wt % has a heat of hydrogen
absorption of ca. 33.7 kJ/mol-H2 which is in excellent agreement with the value (33.6 kJ/mol-H2) predicted theoretically. The
partially reversible hydrogen release and uptake can be can be achieved at the temperature as low as 100 °C. These results
illustrate the great potential of metalorganic hydrides for tackling the grand challenge of hydrogen storage and manifest the
effectiveness of theoretical prediction in guiding materials fabrication.

With the rapid growth of population and economy, the world capacity (≥5.0 wt %) and suitable thermodynamic properties
is facing an energy shortage and environmental pollution. The (the heat of hydrogen desorption (ΔHd) is around 30 kJ/mol-
utilization of renewable and clean energy is an urgent task of H2). However, it is extremely difficult to meet both criteria
recent research.1 Hydrogen has long been considered as a simultaneously with one material. For instance, sorbents with
clean and efficient energy carrier due to its abundance, high surface area can store more than 5.0 wt % hydrogen
environmental friendliness, high energy density, and recycla- through physisorption; however, they have low ΔHd (0−10 kJ/
bility.2−6 However, one of the bottlenecks and challenges in
mol-H2) which results in low operating temperature.7 Complex
the implementation of hydrogen energy is the lack of efficient
hydrides and LOHCs with high hydrogen densities dehydro-
hydrogen storage method.7,8 Tremendous research efforts have
been made in the past half-century on hydrogen storage,
including physical (such as compression,9 liquefaction,10 and Received: August 11, 2021
physisorption11−14) and chemical methods (such as metal Accepted: August 18, 2021
hydrides,15,16 complex hydrides,17−20 chemical hydrides,21−24
and liquid organic hydrogen carriers (LOHCs)25−28). To
uptake and release a large amount of hydrogen reversibly under
mild conditions, two criteria are to be met, i.e., high hydrogen

© XXXX American Chemical Society https://doi.org/10.1021/acsmaterialslett.1c00488


1417 ACS Materials Lett. 2021, 3, 1417−1425
ACS Materials Letters www.acsmaterialsletters.org Letter

Figure 1. Selected organic compounds having “reactive protic H” for metal substitution.

genate via breaking strong N−H, B−H, and C−H bonds, for the first time, among which 20 pairs show both high
leading to strong endothermicities with high ΔHd (>40 kJ/ hydrogen capacities (≥5 wt %) and suitable thermodynamics
mol-H2).17,18,25,29−32 Several transition metal hydrides includ- (25−35 kJ/mol-H2) for hydrogen storage, showing the
ing LaNi5H6, TiV2H4, and FeTiH2, etc., on the other hand, efficiency of theoretical prediction and great potential of
possess suitable thermodynamics and high volumetric hydro- metalorganic hydrides. Based on the calculations, four of them
gen density, unfortunately, their gravimetric hydrogen densities are selected as examples to be synthesized, characterized, and
are usually lower than 3.0 wt %.15,33 Therefore, the investigated in detail. Specifically, lithium indolide/lithium
development of new materials with both ideal thermodynamics octahydroindolide pair with a theoretical hydrogen capacity of
and high hydrogen capacities is a great challenge. 6.1 wt % and an ideal ΔHd of 33.7 kJ/mol-H2 achieves partially
Very recently, metalorganic hydrides, a new family of reversible hydrogen uptake and release at the temperature as
hydrogen storage materials spanning across the domain of low as 100 °C. The successfulness of theoretical prediction to
inorganic and organic hydrides, were developed.34,35 Thanks to rational fabrication shown in this paper provides an efficient
the electron-donating ability of alkali or alkaline earth metal, and inspiring method to develop promising materials for
the thermodynamics of metalorganic hydrides could be hydrogen storage.
rationally modulated. The versatile chemistry of organics and Theoretical Screen. The strategy employed here is the
metal provides a vast opportunity for discovering new and metal substitution of reactive H in organic compounds forming
promising materials for hydrogen storage. However, every coin metalorganic hydrides to modulate the dehydrogenation
has two sides. The variable properties of organics and metals thermodynamics. As shown in Figure 1, selected organic
inevitably complicate the material development. Screening compounds employed as potential candidates for the metal
promising materials from a larger number of candidates via substitution include carbocyclic or heterocyclic molecules
routine experimental syntheses and property evaluations is having “reactive protic H” that bonds to N, O, or S within or
extremely inefficient. Until now, only a couple of candidates out of the organic ring. All the organic compounds selected are
which satisfy the two aforementioned key criteria were commercially available.
developed, such as lithium and sodium carbazolides.36 Light alkali metals, Li and Na, are employed to replace the
Therefore, it is highly demanding to screen materials by reactive H in the selected 47 organic compounds (Figure 1) to
wide-ranging theoretical prediction, aiming to guide exper- form 94 metalorganic compounds to be screened in the
imental synthesis and improve efficiency. In the present study, present study. The X1s method37 was used to calculate the
more than 90 metalorganic hydrides are screened theoretically hydrogenation enthalpy change (ΔHh) of the H-lean to H-rich
1418 https://doi.org/10.1021/acsmaterialslett.1c00488
ACS Materials Lett. 2021, 3, 1417−1425
ACS Materials Letters www.acsmaterialsletters.org Letter

Figure 2. (a) Hydrogen storage properties (ΔHd and gravimetric hydrogen density) of metalorganic hydrides (gray for the pristine organics,
blue and red for lithium- and sodium-substituted compounds, respectively) compared with representative materials (green). (b) ΔHd of 8H-
M-ID and 8H-M-AD versus electronegativities of H and metals. (c) The lengths of C−H bonds at the α site of 8H-M-ID and 8H-M-AD
versus electronegativities of H and metals.

metalorganic hydrides compared with their parent compounds As mentioned in the introduction part, promising hydrogen
as shown in Table S1. Note that the absolute value of ΔHh storage materials should meet two criteria, i.e., suitable
equals to the ΔHd of H-rich counterpart. It is shown that the thermodynamics and high hydrogen capacity. To identify
ΔHd of organic compounds can be efficiently reduced through potential candidates, the calculated ΔHd of metalorganic
metalation, which is consistent with the previous reports.34,36,38 hydrides and their hydrogen capacities are summarized in
The reason can be ascribed to the fact that both H-lean and H- Figure 2a. It is obvious that these parent compounds (gray)
rich compounds are stabilized upon metalation. However, the move to low ΔHd region upon metalation (blue and red) with
H-lean counterparts are more stabilized than the H-rich ones slightly scarifying hydrogen capacities. Theoretically, to achieve
due to the fact that the magnitude of the conjugation effect is reversible hydrogen uptake and release at moderate conditions,
significantly greater than that of the hyper-conjugation effect; ΔGr for one reaction should be around zero. Assuming ΔSd is
i.e., the electrons transferred from the metal to the aromatic around 100−120 J/(K mol-H2), an ideal area for hydrogen
storage (25 kJ/mol-H2 ≤ ΔHd ≤ 35 kJ/mol-H2 and hydrogen
ring in H-lean compounds are more easily delocalized than
capacity ≥5.0 wt %) is marked in orange in the figure. Barely
those of H-rich compounds, resulting in more stabilized H-lean
any reversible hydrogen storage materials developed to date
counterparts as well as a substantial decrease in ΔHd upon
could meet both criteria at the same time (green). Importantly,
metalation. Moreover, the metal substitution effect in the based on our calculations about 20 metalorganic hydrides, i.e.,
organic compounds that have reactive H on the organic ring is lithiated and/or sodiated 14, 22, 23, 26, 27, 31, 32, 34, 39, 44,
more significant than those with the reactive H out of the 45, 46, and 47, exhibit both high gravimetric hydrogen
organic ring. For instance, the ΔHh of 22, 23, 27, 28, and 29 capacities and suitable ΔHd and sit within this area (red data in
having reactive H on the ring are changed more significantly Table S1), holding great promise for hydrogen storage
upon metal substitution than those of 25, 26, and 30 that have application as compared with representative LOHCs (such as
reactive H out of the organic ring. It is worth mentioning that 12H-NEC and MCH). Comparatively, metalorganic hydrides
hydrogenation of H-lean metalorganic hydride could result in with fused-ring structures exhibit superior properties to their
the decrease of ΔHh in some cases especially the ones with S in single-ring counterparts (such as 3, 4, 7, 9, etc.), which may be
the ring (data in blue in Table S1), which is ascribed to the due to the facts that 1) the fused-ring structure could average
ring-opening reaction during hydrogenation. the electron donation effect of metal and avoid over
1419 https://doi.org/10.1021/acsmaterialslett.1c00488
ACS Materials Lett. 2021, 3, 1417−1425
ACS Materials Letters www.acsmaterialsletters.org Letter

Figure 3. Simulated photoelectron spectra (black lines) of the lowest energy isomer for [(indole)2·H]− (a) and [(indole)2·M]− (M = Li, Na,
K) (b−d) in comparison with the corresponding experimental spectra (red lines). The calculated stick spectrum for each complex was
generated using the measured VDE and TDDFT excitation energies. The simulated spectrum was obtained by convoluting each stick with
Gaussian functions of 0.03 eV fwhm in accord with the experimental conditions. In simulations the binding energies are explicitly
considered, but with the assumption of uniform peak intensity. (e) Electron localization function of [(indole)2·H]− and [(indole)2·M]− (M =
Li, Na, K).

Table 1. Comparison of Experimental VDE and ADE Values to the Calculated Ones of Lowest-Energy Isomers for [(indole)2·
H]− and [(indole)2·M]− (M = Li, Na, K)a
complex [(indole)2·H]− [(indole)2·Li]− [(indole)2·Na]− [(indole)2·K]−
expt VDE (eV) 3.26 3.83 3.80 3.83
expt ADE (eV) 3.13 3.81 3.77 3.72
calc VDE (eV) 3.27 3.66/3.66 3.61/3.62 3.52/3.52
calc ADE (eV) 3.03 3.58/3.58 3.53/3.54 3.44/3.44
natural charge of H/Li/Na/K (e) 0.476 0.834/0.831 0.884/0.886 0.923/0.922
natural charge of indole ring (e) −1.476 −1.834/-1.831 −1.884/-1.886 −1.923/-1.922
a
The natural charge of the metal and indole ring are also listed.

optimization of ΔHd, and 2) the fused-ring has more C−H be tailored via changing metal. For instance, the ΔHd of 8H-Li-
bonds and larger molecular weight that could lessen the ID and 8H-Na-ID decrease significantly from that of 8H-ID
scarifying effect of the weight of metal elements (Li or Na) on (53.0 kJ/mol-H2) to suitable values of 33.7 and 30.8 kJ/mol-
the hydrogen capacity. In the following sections, a few H2, respectively. This is because both H-lean and H-rich
examples selected from the marked area will be investigated compounds are stabilized upon metalation, while the
in detail. conjugation effect in H-lean compounds is significantly greater
Thermodynamics vs Electron Properties. Among these than the hyper-conjugation effect in H-rich compounds; thus
materials in the ideal area, metal indolide/octahydroindolide the metalation causes a stronger stabilization for the H-lean
(22, M-ID/8H-M-ID for short) and metal 7-azaindolide/ compounds resulting in a decrease in ΔHd.34,36,38 It is also
octahydro-7-azaindolide (23, M-AD/8H-M-AD for short) found that the lengths of C−H bond at the α-site of 8H-M-ID
pairs were selected for further investigation due to their low- and 8H-M-AD increase upon metalation; more importantly,
cost, high hydrogen capacities, suitable thermodynamics, and they increase as the decrease of electronegativity of metal as
ease of syntheses. To figure out the relationship between shown in Figure 2c, manifesting that the C−H bond at α-site
hydrogen storage thermodynamic properties and electron can be activated to a certain extent such that it is likely to be
properties of these metalorganic hydrides, the ΔHd of M-ID/ dissociated first upon dehydrogenation.
8H-M-ID pairs and M-AD/8H-M-AD pairs (M = Li, Na, K, To further provide direct experimental evidence that metal
Mg, Ca) are calculated using the X1s method (Figure 2b). It is substitution can stabilize the electron of indole and azaindole
shown that the ΔHd of 8H-M-ID and 8H-M-AD decreases rings, the negatively charged complexes consisting of two
with the Pauling electronegativity (χ) of metals, exhibiting a indole/azaindole rings and one cation (H+, Li+, Na+, K+) were
nearly linear relationship and manifesting that the ΔHd could studied in the gas phase using anion photoelectron spectros-
1420 https://doi.org/10.1021/acsmaterialslett.1c00488
ACS Materials Lett. 2021, 3, 1417−1425
ACS Materials Letters www.acsmaterialsletters.org Letter

Figure 4. (a and b) 1H and 13C NMR spectra of synthesized M-ID and M-AD compared to their parent organic compounds in DMSO-d6. (c)
XRD patterns of synthesized M-ID and M-AD compared to pristine reagents. Crystal structures (left) and local coordination of cations
(right) of (d) Li-ID and (e) Na-AD. Color: blue sphere, nitrogen; yellow sphere, lithium; green sphere, sodium; black sphere, carbon; pink
sphere, hydrogen.

copy (PES) and quantum chemical calculations. [(indole)2· metal, more importantly, as the increase of the electron-
X]− and [(azaindole)2·X]− (X = H, Li, Na, K) anions are donating ability of alkali metal, the electron density in indole
produced by electrospray ionization method and characterized ring increases, agreeing well with the discussion above. The
by PES (see Supporting Information for more details). As PES and quantum chemical calculations results of [(azain-
shown in Figure 3, two relatively broad spectral bands in the dole)2·H]− and [(azaindole)2·M]− (M = Li, Na, K) are shown
binding energy range of 3.0−4.3 eV can be seen in [(indole)2· in Figure S2, Table S2, and Table S3, which show a similar
H]−. In contrast, the [(indole)2·M]− anions exhibit narrower tendency to that of M-ID.
peaks with their electron binding energies ∼0.7 eV higher Material Syntheses, Characterizations, and Crystal
compared to [(indole)2·H]−, showing metal substitution Structures. Two methods, i.e., wet chemical method and
stabilizes the organic structure, consistent with the discussion heat-treatment method, were used to synthesize M-ID and M-
above. Theoretical calculations indicated the optimized AD by reacting metal hydrides (lithium hydride and sodium
structure of [(indole)2·H]− with two indole rings perpendic- hydride) and pristine indole and 7-azaindole via reactions 1
ular to each other along the N−H−N axis, while each metal and 2, respectively. Around 1 equiv of hydrogen can be
substitution cluster has two close-lying structures with the two released from all the reactions as shown in Figures S3−S5,
indole rings in the same planar but with different directions indicating the formation of corresponding metalorganic
(Figure 3, Table S3). The calculated vertical and adiabatic hydrides.
detachment energies (VDEs and ADEs) (Table 1) and
simulated spectra (Figure 3) compare reasonably well with
experimental data in term of the spectral pattern and peak
position (see Supporting Information for the details of
simulations). The congested spectral band in [(indole)2·H]−
(Figure 3a) and an extra peak in [(indole)2·Li]− (Figure 3b),
compared to the their corresponding simulated ones, is most
likely due to vibrational excitations that are not considered in
the simulations. Since both isomers of [(indole)2·M]− show X-ray diffraction (XRD) and nuclear magnetic resonance
close spectral bands and VDEs compared to the experiments, it (NMR) were used to characterize the synthesized compounds
is likely that they are both formed and coexist in the (Li/Na-ID and Li/Na-AD). As shown in 1H NMR spectra
experiments. Figure 3e shows the calculated renderings of (Figure 4a), the 1H signals of N−H at 11.09 and 11.62 ppm in
the electron localization function surface using the optimized indole and 7-azaindole disappear in all of the synthesized
structures indicating dominant electrostatic interactions metalorganic hydrides, showing that the reactive H on N−H
between M and indole rings. The natural charge analyses was replaced by metal. In addition, 1H signals of C−H move to
(Table 1) provide an apparent indication that the charge the high field after the metal substitution, indicating the
density of indole ring increased with the introduction of alkali enhancement of the shielding effect due to the increase of
1421 https://doi.org/10.1021/acsmaterialslett.1c00488
ACS Materials Lett. 2021, 3, 1417−1425
ACS Materials Letters www.acsmaterialsletters.org Letter

Figure 5. (a) TEM images of in situ synthesized Ru nanoparticles. The inset figure is the magnified image of Ru nanoparticles. (b)
Conversion of hydrogenation of Li-ID to 8H-Li-ID catalyzed by Ru nanoparticles. (c) 1H NMR spectra of hydrogenated Li-ID catalyzed by
Ru nanoparticles and dehydrogenated Li-8H-ID catalyzed by Pt nanoparticles in DMSO-d6. (d) Microcalorimetry measurements of
hydrogenation of Li-ID catalyzed by in situ formed Ru nanoparticles.

electron density on the organic ring. Moreover, the chemical interaction and directly contacting with Ns of the other two
shifts of sodium-substituted organic compounds are larger than indolides via cation−N σ bonding. The indolide fragments are
that of lithium reflecting the stronger electron-donating effect thus linked together by Li+ cations through alternating cation−
of sodium, which is consistent with our theoretical calculation N σ bonding and cation−π interaction, forming a 1-D chain in
results. Nevertheless, the 13C NMR signals move to different a herringbone fashion along the b-axis of the unit cell. The
directions upon metalation (Figure 4b). For instance, the adjacent chains are interacting each other through hydrogen
signals of C atoms at 1, 3, 7, and 8 sites move to the low field, bonding, which helps construct a 3-D structure. For Na-AD,
and the ones at 2, 4, 5, and 6 sites move to the high field in Li- each Na+ cation connects with four surrounding azaindolides
ID (Figure S6). This is attributed to the charge redistribution through cation−π interaction with one azaindolide, and σ−N
due to the conjugation effect on the organic ring, which is also bonding with the Ns from the pyrrole rings of two anions, and
demonstrated by the merging of several 1H signals in Na/Li-ID with the N from the pyridine ring of the fourth azaindolide.
(Figure S6). Furthermore, the appearance of strong 7Li and The azaindolide anions are knitted together by sodium cations
23
Na NMR signals of M-ID and M-AD in DMSO-d6 suggests and woven into a hollow cage-like cluster extending infinitely
the formation of soluble metal salts (Figure S8). The along the c-axis of the unit cell. Such extended cage-like
diffraction patterns of the synthesized metalorganic hydrides clusters interact with the adjacent ones through hydrogen
are different from those of the pristine metal hydrides and bonding, forming a stable 3-D structure. It is worth noting that
organic compounds (Figure 4c), indicating the formation of the planar structures of the gaseous [(indole)2·M]− and
new compounds. [(azaindole)2·M]− (M = Li, Na, K) anions, mentioned in
The crystal structures of Li-ID and Na-AD were determined Figures 3 and S2, are not found in the local structures of the
using powder XRD, optimized by DFT calculations, and crystals, which may be ascribed to the facts that 1) cation−N σ
refined with the GSAS package (see Supporting Information bonding is usually stronger than that of cation−π interaction,
for more details). Li-ID and Na-AD crystallize in monoclinic preferably leading to formation of cation−N bonding in the
P21/c (no. 14) and tetragonal P4̅21c (no. 114) space groups planar anions that are calculated from gas phase simulation,
with lattice parameters of a = 9.2902(5) Å, b = 5.1956(4) Å, c and 2) the 3-D structures in the crystals allow steric interaction
= 13.7739(16) Å, β = 94.759(4)°, V = 662.55(16) Å3, and a = between cation and anion, resulting in both cation−N σ
13.5659(6) Å, c = 7.1869(4) Å, V = 1322.64(15) Å3, bonding and cation−π interaction. The existence of cation−π
respectively (Figures S9 and S10). The crystal structures of interaction in these two structures inevitably stabilizes the H-
Li-ID and Na-AD and the local coordination of the Li+/Na+ lean counterparts, which will further optimize the energy gap
cations are shown in Figures 4d and 4e, respectively. For Li-ID, between H-lean and H-rich metalorganic hydrides. These
each Li+ cation is interacting with three surrounding indolides crystal structures provide insightful details about cation−anion
by facing the aromatic ring of one indolide via cation−π and cation−π arrangement, which will conversely help
1422 https://doi.org/10.1021/acsmaterialslett.1c00488
ACS Materials Lett. 2021, 3, 1417−1425
ACS Materials Letters www.acsmaterialsletters.org Letter

modulate their thermodynamic parameters from DFT usually high than 65.0 kJ/mol-H2); the thermal dehydrogen-
calculations. ation temperature is higher than 300 °C.42,43 For the well-
Hydrogen Storage Properties. Guided by the thermody- developed N-heterocycles (such as 12H-NEC), the thermal
namic prediction, experiments on reversible hydrogen release dehydrogenation temperatures are also higher than 170
and uptake are conducted. Among these four synthesized °C.44,45
metalorganic hydrides (Li/Na-ID and Li/Na-AD), Li-ID with With an ideal thermodynamic property and high hydrogen
a hydrogen storage capacity of 6.1 wt % and theoretical ΔHh of content, the Li-ID/8H-Li-ID pair would facilitate the ease of
−33.7 kJ/mol-H2 as shown in reaction 3 is selected and
hydrogen uptake and release in the present study. Considering
investigated for hydrogen storage. The as-prepared Li-ID is
hard to be hydrogenated under 50 bar hydrogen at 100 °C that indole and 7-azaindole, the precursors of metalorganic
without catalyst likely due to the presence of high kinetic hydrides, are low-cost chemicals with high annual production,
barrier since catalysts are always needed in the hydrogenation together with their facile syntheses, high hydrogen capacities,
and dehydrogenation of LOCHs.25,39,40 Therefore, Ru and Pt and suitable thermodynamic properties, M-ID and M-AD hold
nanoparticles were in situ synthesized as the hydrogenation great promise for both on-board application and large-scale
and dehydrogenation catalysts, respectively (see Supporting long-distance hydrogen transportation. In addition to them,
Information for more detail). Ru nanoparticles with an other potential candidates predicted with high hydrogen
averaging particle size of 2.1 nm are well dispersed in Li-ID capacity and suitable thermodynamic properties in the ideal
sample (Figures 5 and S11−S13). It is shown that the area (Figure 2) are worthy of investigation in the future.
hydrogenation of Li-ID can be achieved under 50 bar Theoretical calculations were used to screen promising
hydrogen at 100 °C in the presence of Ru nanoparticles metalorganic hydrides for hydrogen storage. More than 90
(Figure 5b). Conversion of Li-ID higher than 99% was metalorganic hydrides were calculated, among which 20
obtained in 15 min with selectivity to 8H-Li-ID higher than candidates with high hydrogen storage capacity (≥5.0 wt %)
99% as evidenced by the quantitative 1H NMR analysis (Figure
and suitable thermodynamic (ΔHd of 25−35 kJ/mol-H2) were
5c). Since Li-ID can be almost totally converted to 8H-Li-ID
identified. It is found that the ΔHd of metalorganic hydrides
(Figure 5b), the hydrogenation enthalpy change (ΔHh) was
measured using a C80 calorimeter. The ΔHh of Li-AD can be tailored by the electronegativity of the metal. Four of
determined experimentally is −33.6 kJ/mol-H2 (Figure 5d), them, i.e., Li/Na-ID and Li/Na-AD, are selected for
which is in excellent agreement with the calculated value (33.7 experimental validation. All the four candidates can be easily
kJ/mol-H2 for ΔHd), validating the accuracy of our simulation synthesized via the reactions between metal hydrides and
(see Figure S14 for more details). corresponding organic precursors. Among them, the structures
of lithium indolide (P21/c (no. 14)) and sodium 7-azaindolide
(P4̅21c (no. 114) were successfully resolved. PES and quantum
chemical calculations confirmed that the charge density of the
organic ring increased with the introduction of the alkali metal,
In the attempt of dehydrogenation of 8H-Li-ID, we used Pt thus optimizing their ΔHd for hydrogen storage. Importantly,
nanoparticles as catalysts because Ru nanoparticles are inactive the partially reversible hydrogenation and dehydrogenation of
to catalyze the dehydrogenation. It is shown that around 51.4% Li-ID/8H-Li-ID pair with an ideal ΔHd of 33.7 kJ/mol-H2
8H-Li-ID converts to Li-ID under vacuum at 100 °C in the were successfully achieved at the temperature as low as 100 °C
presence of Pt nanoparticles (Figure 5c), indicating 3.1 wt % in the presence of Ru and Pt nanoparticles. The theoretical
hydrogen is reversibly stored. Because the conversion to Li-ID hydrogen storage capacity of this material is as high as 6.1 wt
remained constant even after long dehydrogenation time, we %. The development of efficient catalysts for hydrogenation
suppose that Pt nanoparticles were covered or separated by Li-
and dehydrogenation is highly needed to further optimize the
ID from contacting 8H-Li-ID, leading to mass transfer
constrain which is a common issue in solid-state catalysis. kinetics. In addition, other potential candidates predicted with
The development of efficient catalysts or optimization of high hydrogen capacity and suitable thermodynamic properties
dehydrogenation process are thus very important. Catalysis in in the ideal area are worthy of investigation in the future.
solution may overcome this mass transfer issue; however, it These results illustrated that the strategy from theoretical
may sacrifice the hydrogen capacity. There are several good prediction to rational fabrication is successful and represents
examples of solid-state catalysis in hydrogen storage area, such an efficient methodology to develop promising materials for
as Ti-based catalysts for NaAlH417 and K-based catalyst for hydrogen storage.


amide-hydride system,41 which highlight unique function of
nonconventional catalysts. Nevertheless, we managed to ASSOCIATED CONTENT
demonstrate a reversible hydrogenation and dehydrogenation
of Li-ID/8H-Li-ID pair at a temperature as low as 100 °C in *
sı Supporting Information

the presence of Ru and Pt nanoparticles, albeit the conversion The Supporting Information is available free of charge at
and kinetics still need to be optimized. It should be noted that https://pubs.acs.org/doi/10.1021/acsmaterialslett.1c00488.
no conversion of pristine octahydroindole to indole can be
obtained under the same conditions in the presence of the Experimental procedures, Table S1−S3 (enthalpies and
same catalyst. It is also worth mentioning that the dehydrogen- hydrogen capacities, comparison of VDE and ADE
ation of conventional LOHCs usually needs much higher values, relative energies), and Figure S1−S14 (TPD-MS,
temperatures. For instance, the dehydrogenation of cyclo- photoelectron spectra, hydrogen evolution, NMR
alkanes (such as methylcyclohexane and dodecahydro- spectra, XRD profiles, TEM images and particle size
benzyltoluene) is a strong endothermic process (ΔHd is distribution, EDS, microcalorimetry) (PDF)

1423 https://doi.org/10.1021/acsmaterialslett.1c00488
ACS Materials Lett. 2021, 3, 1417−1425
ACS Materials Letters


www.acsmaterialsletters.org Letter

AUTHOR INFORMATION (XLYC1807157), the K. C. Wong Education Foundation


Corresponding Authors (GJTD-2018-06), and the Dalian high-level talents program
Xue-Bin Wang − Physical Sciences Division, Pacific Northwest (2019RD09). We also are thankful for beam time from
National Laboratory, Richland, Washington 99352, United BL14B1 at SSRF for the XRD experiment. The PES work was
States; orcid.org/0000-0001-8326-1780; supported by the U.S. Department of Energy (DOE), Office of
Email: xuebin.wang@pnnl.gov Science, Office of Basic Energy Sciences, Division of Chemical
Hui Wu − NIST Center for Neutron Research, National Sciences, Geosciences, and Biosciences, and performed using
Institute of Standards and Technology, Gaithersburg, EMSL, a national scientific user facility sponsored by DOE’s
Maryland 20899-6102, United States; orcid.org/0000- Office of Biological and Environmental Research and located at
0003-0296-5204; Email: huiwu@nist.gov Pacific Northwest National Laboratory, which is operated by
Anan Wu − Fujian Provincial Key Laboratory of Theoretical Battelle Memorial Institute for the DOE.
and Computational Chemistry, College of Chemistry and
Chemical Engineering, Xiamen University, Xiamen 361005,
China; orcid.org/0000-0001-5243-9291;
■ REFERENCES
(1) Nejat, P.; Jomehzadeh, F.; Taheri, M. M.; Gohari, M.; Abd.
Email: ananwu@xmu.edu.cn Majid, M. Z. A global review of energy consumption, CO2 emissions
and policy in the residential sector (with an overview of the top ten
Teng He − Dalian Institute of Chemical Physics, Chinese
CO2 emitting countries). Renewable Sustainable Energy Rev. 2015, 43,
Academy of Sciences, Dalian 116023, China; orcid.org/ 843−862.
0000-0003-2900-7612; Email: heteng@dicp.ac.cn (2) Schlapbach, L.; Züttel, A. Hydrogen-storage materials for mobile
applications. Nature 2001, 414, 353−358.
Authors (3) Johnston, B.; Mayo, M. C.; Khare, A. Hydrogen: the energy
Zijun Jing − Dalian Institute of Chemical Physics, Chinese source for the 21st century. Technovation 2005, 25, 569−585.
Academy of Sciences, Dalian 116023, China; University of (4) Ball, M.; Wietschel, M. The future of hydrogen−opportunities
Chinese Academy of Sciences, Beijing 100049, China and challenges. Int. J. Hydrogen Energy 2009, 34, 615−627.
Qinqin Yuan − Physical Sciences Division, Pacific Northwest (5) Abe, J. O.; Popoola, A. P. I.; Ajenifuja, E.; Popoola, O. M.
National Laboratory, Richland, Washington 99352, United Hydrogen energy, economy and storage: Review and recommenda-
States; orcid.org/0000-0001-5771-6147 tion. Int. J. Hydrogen Energy 2019, 44, 15072−15086.
Yang Yu − Dalian Institute of Chemical Physics, Chinese (6) Abdalla, A. M.; Hossain, S.; Nisfindy, O. B.; Azad, A. T.;
Academy of Sciences, Dalian 116023, China Dawood, M.; Azad, A. K. Hydrogen production, storage, trans-
portation and key challenges with applications: A review. Energy
Xiangtao Kong − College of Chemistry and Chemical Convers. Manage. 2018, 165, 602−627.
Engineering, Anyang Normal University, Anyang 455000, P. (7) He, T.; Pachfule, P.; Wu, H.; Xu, Q.; Chen, P. Hydrogen
R. China carriers. Nat. Rev. Mater. 2016, 1, 16059.
Khai Chen Tan − Dalian Institute of Chemical Physics, (8) Mohtadi, R.; Orimo, S.-i. The renaissance of hydrides as energy
Chinese Academy of Sciences, Dalian 116023, China; materials. Nat. Rev. Mater. 2017, 2, 16091.
University of Chinese Academy of Sciences, Beijing 100049, (9) Zheng, J.; Liu, X.; Xu, P.; Liu, P.; Zhao, Y.; Yang, J. Development
China of high pressure gaseous hydrogen storage technologies. Int. J.
Jintao Wang − Dalian Institute of Chemical Physics, Chinese Hydrogen Energy 2012, 37, 1048−1057.
Academy of Sciences, Dalian 116023, China; University of (10) Peschka, W.; Carpetis, C. Cryogenic hydrogen storage and
Chinese Academy of Sciences, Beijing 100049, China refueling for automobiles. Int. J. Hydrogen Energy 1980, 5, 619−625.
(11) Dillon, A. C.; Jones, K. M.; Bekkedahl, T. A.; Kiang, C. H.;
Qijun Pei − Dalian Institute of Chemical Physics, Chinese
Bethune, D. S.; Heben, M. J. Storage of hydrogen in single-walled
Academy of Sciences, Dalian 116023, China carbon nanotubes. Nature 1997, 386, 377−379.
Wei Zhou − NIST Center for Neutron Research, National (12) Rosi, N. L.; Eckert, J.; Eddaoudi, M.; Vodak, D. T.; Kim, J.;
Institute of Standards and Technology, Gaithersburg, Keeffe, M.; Yaghi, O. M. Hydrogen Storage in Microporous Metal-
Maryland 20899-6102, United States; orcid.org/0000- Organic Frameworks. Science 2003, 300, 1127.
0002-5461-3617 (13) Chae, H. K.; Siberio-Pérez, D. Y.; Kim, J.; Go, Y.; Eddaoudi,
Ping Chen − Dalian Institute of Chemical Physics, Chinese M.; Matzger, A. J.; O’Keeffe, M.; Yaghi, O. M.; Materials, D.;
Academy of Sciences, Dalian 116023, China; orcid.org/ Discovery, G. A route to high surface area, porosity and inclusion of
0000-0002-0625-0639 large molecules in crystals. Nature 2004, 427, 523−527.
(14) Liu, S.; Liu, J.; Liu, X.; Shang, J.; Xu, L.; Yu, R.; Shui, J.
Complete contact information is available at: Hydrogen storage in incompletely etched multilayer Ti2CTx at room
https://pubs.acs.org/10.1021/acsmaterialslett.1c00488 temperature. Nat. Nanotechnol. 2021, 16, 331−336.
(15) Sakintuna, B.; Lamari-Darkrim, F.; Hirscher, M. Metal hydride
Author Contributions materials for solid hydrogen storage: A review. Int. J. Hydrogen Energy
The manuscript was written through contributions of all 2007, 32, 1121−1140.
authors. All authors have given approval to the final version of (16) Zhang, X.; Liu, Y.; Ren, Z.; Zhang, X.; Hu, J.; Huang, Z.; Lu, Y.;
the manuscript. Gao, M.; Pan, H. Realizing 6.7 wt% reversible storage of hydrogen at
ambient temperature with non-confined ultrafine magnesium
Notes hydrides. Energy Environ. Sci. 2021, 14, 2302−2313.
The authors declare no competing financial interest. (17) Bogdanovic, B.; Schwickardi, M. Ti-doped alkali metal


aluminium hydrides as potential novel reversible hydrogen storage
ACKNOWLEDGMENTS materials. J. Alloys Compd. 1997, 253, 1−9.
(18) Chen, P.; Xiong, Z.; Luo, J.; Lin, J.; Tan, K. L. Interaction of
This work is partially supported by the National Key R&D hydrogen with metal nitrides and imides. Nature 2002, 420, 302−304.
Program of China (2019YFE0103600), the National Natural (19) Züttel, A.; Wenger, P.; Rentsch, S.; Sudan, P.; Mauron, P.;
Science Foundation of China (21875246, 21988101, Emmenegger, C. LiBH4 a new hydrogen storage material. J. Power
21773193), the LiaoNing Revitalization Talents Program Sources 2003, 118, 1−7.

1424 https://doi.org/10.1021/acsmaterialslett.1c00488
ACS Materials Lett. 2021, 3, 1417−1425
ACS Materials Letters www.acsmaterialsletters.org Letter

(20) Hirscher, M.; Yartys, V. A.; Baricco, M.; Bellosta von Colbe, J.; (39) Aakko-Saksa, P. T.; Cook, C.; Kiviaho, J.; Repo, T. Liquid
Blanchard, D.; Bowman, R. C.; Broom, D. P.; Buckley, C. E.; Chang, organic hydrogen carriers for transportation and storing of renewable
F.; Chen, P.; Cho, Y. W.; Crivello, J.-C.; Cuevas, F.; David, W. I. F.; energy−Review and discussion. J. Power Sources 2018, 396, 803−823.
de Jongh, P. E.; Denys, R. V.; Dornheim, M.; Felderhoff, M.; (40) Modisha, P. M.; Ouma, C. N. M.; Garidzirai, R.; Wasserscheid,
Filinchuk, Y.; Froudakis, G. E.; Grant, D. M.; Gray, E. M.; Hauback, P.; Bessarabov, D. The Prospect of Hydrogen Storage Using Liquid
B. C.; He, T.; Humphries, T. D.; Jensen, T. R.; Kim, S.; Kojima, Y.; Organic Hydrogen Carriers. Energy Fuels 2019, 33, 2778−2796.
Latroche, M.; Li, H.-W.; Lototskyy, M. V.; Makepeace, J. W.; Møller, (41) Wang, J.; Liu, T.; Wu, G.; Li, W.; Liu, Y.; Araújo, C. M.;
K. T.; Naheed, L.; Ngene, P.; Noréus, D.; Nygård, M. M.; Orimo, S.- Scheicher, R. H.; Blomqvist, A.; Ahuja, R.; Xiong, Z.; Yang, P.; Gao,
i.; Paskevicius, M.; Pasquini, L.; Ravnsbæk, D. B.; Veronica Sofianos, M.; Pan, H.; Chen, P. Potassium-Modified Mg(NH2)2/2 LiH System
M.; Udovic, T. J.; Vegge, T.; Walker, G. S.; Webb, C. J.; Weidenthaler, for Hydrogen Storage. Angew. Chem., Int. Ed. 2009, 48, 5828−5832.
C.; Zlotea, C. Materials for hydrogen-based energy storage−past, (42) Cui, Y.; Kwok, S.; Bucholtz, A.; Davis, B.; Whitney, R. A.;
recent progress and future outlook. J. Alloys Compd. 2020, 827, Jessop, P. G. The effect of substitution on the utility of piperidines
153548. and octahydroindoles for reversible hydrogen storage. New J. Chem.
(21) Stephens, F. H.; Pons, V.; Tom Baker, R. Ammonia−borane: 2008, 32, 1027−1037.
the hydrogen source par excellence? Dalton Trans. 2007, 2613−2626. (43) Biniwale, R. B.; Rayalu, S.; Devotta, S.; Ichikawa, M. Chemical
(22) Singh, S. K.; Singh, A. K.; Aranishi, K.; Xu, Q. Noble-Metal- hydrides: A solution to high capacity hydrogen storage and supply.
Free Bimetallic Nanoparticle-Catalyzed Selective Hydrogen Gener- Int. J. Hydrogen Energy 2008, 33, 360−365.
ation from Hydrous Hydrazine for Chemical Hydrogen Storage. J. (44) Wang, Z.; Tonks, I.; Belli, J.; Jensen, C. M. Dehydrogenation of
Am. Chem. Soc. 2011, 133, 19638−19641. N-ethyl perhydrocarbazole catalyzed by PCP pincer iridium
(23) Gutowska, A.; Li, L.; Shin, Y.; Wang, C. M.; Li, X. S.; Linehan, complexes: Evaluation of a homogenous hydrogen storage system. J.
J. C.; Smith, R. S.; Kay, B. D.; Schmid, B.; Shaw, W.; Gutowski, M.; Organomet. Chem. 2009, 694, 2854−2857.
Autrey, T. Nanoscaffold Mediates Hydrogen Release and the (45) Yang, M.; Han, C.; Ni, G.; Wu, J.; Cheng, H. Temperature
Reactivity of Ammonia Borane. Angew. Chem., Int. Ed. 2005, 44, controlled three-stage catalytic dehydrogenation and cycle perform-
3578−3582. ance of perhydro-9-ethylcarbazole. Int. J. Hydrogen Energy 2012, 37,
(24) Zhu, Y.; Ouyang, L.; Zhong, H.; Liu, J.; Zhu, M. J. A. C. 12839−12845.
Closing the Loop for Hydrogen Storage: Facile Regeneration of
NaBH4 from its Hydrolytic Product. Angew. Chem. 2020, 132, 8701−
8707.
(25) Preuster, P.; Papp, C.; Wasserscheid, P. Liquid Organic
Hydrogen Carriers (LOHCs): Toward a Hydrogen-free Hydrogen
Economy. Acc. Chem. Res. 2017, 50, 74−85.
(26) Zhu, Q.-L.; Xu, Q. Liquid organic and inorganic chemical
hydrides for high-capacity hydrogen storage. Energy Environ. Sci. 2015,
8, 478−512.
(27) He, T.; Pei, Q.; Chen, P. Liquid organic hydrogen carriers. J.
Energy Chem. 2015, 24, 587−594.
(28) Wang, C.; Astruc, D. Recent developments of nanocatalyzed
liquid-phase hydrogen generation. Chem. Soc. Rev. 2021, 50, 3437−
3484.
(29) Yang, J.; Sudik, A.; Wolverton, C.; Siegel, D. J. High capacity
hydrogen storage materials: attributes for automotive applications and
techniques for materials discovery. Chem. Soc. Rev. 2010, 39, 656−
675.
(30) Yadav, M.; Xu, Q. Liquid-phase chemical hydrogen storage
materials. Energy Environ. Sci. 2012, 5, 9698−9725.
(31) Graetz, J. New approaches to hydrogen storage. Chem. Soc. Rev.
2009, 38, 73−82.
(32) Vajo, J. J.; Skeith, S. L.; Mertens, F. Reversible Storage of
Hydrogen in Destabilized LiBH4. J. Phys. Chem. B 2005, 109, 3719−
3722.
(33) Grochala, W.; Edwards, P. P. Thermal Decomposition of the
Non-Interstitial Hydrides for the Storage and Production of
Hydrogen. Chem. Rev. 2004, 104, 1283−1316.
(34) Yu, Y.; He, T.; Wu, A.; Pei, Q.; Karkamkar, A.; Autrey, T.;
Chen, P. Reversible Hydrogen Uptake/Release over a Sodium
Phenoxide−Cyclohexanolate Pair. Angew. Chem., Int. Ed. 2019, 58,
3102−3107.
(35) He, T.; Cao, H.; Chen, P. Complex Hydrides for Energy
Storage, Conversion, and Utilization. Adv. Mater. 2019, 31, 1902757.
(36) Tan, K. C.; Yu, Y.; Chen, R.; He, T.; Jing, Z.; Pei, Q.; Wang, J.;
Chua, Y. S.; Wu, A.; Zhou, W.; Wu, H.; Chen, P. Metallo-N-
Heterocycles-A new family of hydrogen storage material. Energy
Storage Mater. 2020, 26, 198−202.
(37) Wu, J.; Ying Zhang, I.; Xu, X. The X1s Method for Accurate
Bond Dissociation Energies. ChemPhysChem 2010, 11, 2561−2567.
(38) Jing, Z.; Yu, Y.; Chen, R.; Tan, K. C.; He, T.; Wu, A.; Pei, Q.;
Chua, Y. S.; Zheng, D.; Zhang, X.; Ge, Z.; Zhang, F.; Chen, P. Sodium
anilinide−cyclohexylamide pair: synthesis, characterization, and
hydrogen storage properties. Chem. Commun. 2020, 56, 1944−1947.

1425 https://doi.org/10.1021/acsmaterialslett.1c00488
ACS Materials Lett. 2021, 3, 1417−1425

You might also like