You are on page 1of 8

Research Article

ISSN 1998-0124 CN 11-5974/O4


2021, 14(2): 518–525 https://doi.org/10.1007/s12274-019-2571-9

Highly effective H2/D2 separation in a stable Cu-based metal-


organic framework
Yanan Si1,3, Xiang He2 (), Jie Jiang2, Zhiming Duan2, Wenjing Wang1, and Daqiang Yuan1 ()
1
State Key Laboratory of Structural Chemistry, Fujian Institute of Research on the Structure of Matter, Chinese Academy of Sciences, Fuzhou
350002, China
2
Department of Chemistry, College of Sciences, Shanghai University, Shanghai 200444, China
3
University of the Chinese Academy of Sciences, Beijing 100049, China

© Tsinghua University Press and Springer-Verlag GmbH Germany, part of Springer Nature 2019
Received: 14 October 2019 / Revised: 12 November 2019 / Accepted: 17 November 2019

ABSTRACT
A three-dimensional copper metal–organic framework with the rare chabazite (CHA) topology namely FJI-Y11 has been constructed
with flexibly carboxylic ligand 5,5'-[(1,4-phenylenebis(methylene))bis(oxy)]diisophthalic acid (H4L). FJI-Y11 exhibits high water stability
with the pH range from 2 to 12 at temperature as high as 373 K. Importantly, FJI-Y11 also shows high efficiency of hydrogen isotope
separation using dynamic column breakthrough experiments under atmospheric pressure at 77 K. Attributed to its excellent structural
stability, FJI-Y11 possesses good regenerated performance and maintains high separation efficiency after three cycles of
breakthrough experiments.

KEYWORDS
metal–organic framework (MOFs), hydrogen isotope, separation, breakthrough experiments

diversity, adjustable pore size and high specific surface area


1 Introduction
[27–29]. In particular, the variable pore structures and the
As an essential raw material, deuterium plays an indispensable high-concentration open metal sites of MOFs make them as
role in nuclear fusion reactor [1–4]. Besides, deuterium also ideal materials for hydrogen isotope separation. Recent researches
has a wide range of applications in modern industry, especially about separating hydrogen isotope were mainly focused on
in isotope tracing, neutron scattering and proton nuclear regulating the channel size of MOFs based on KQS or changing
magnetic resonance spectroscopy [5–9]. However, hydrogen the open metal sites interacting with hydrogen isotope based
isotope separation is one of the most challenges in the current on chemical affinity quantum sieve (CAQS) [19, 21, 30–41].
separation technology because of their nearly identical molecular Based on these two mechanisms, some MOFs have been
sizes and physicochemical properties. Nowadays, low-temperature successfully used for improving the separation efficiency of
distillation and heavy water electrolysis are mainly used for hydrogen isotope, such as MFU-4 [31], Cu(I)-MFU-4l, COP-
the separation of hydrogen isotope on industrial plant scale. 27-Co, etc. [33, 42]. Another strategy is integrating KQS and
However, current separation technologies still exist problems CAQS into the same MOFs via the post-modification method.
of high energy consumption and low separation factors [10]. For example, the reaction of completely activated MOF-74-Ni
Recently, porous materials based on physical adsorption for and imidazole molecules can synthesize imidazole partially
separating hydrogen isotope at high temperature have drawn coordinated MOF-74-Ni-IM, which retained partial open
more attention due to their low energy consumption. metal sites and decreased the pore size at the same time [35].
Kinetic quantum sieving (KQS), which was first put forward Although some achievements have been made in MOFs for
by Beenakker et al. in 1995 using a cylindrical pore model, has hydrogen isotope separation, high capacity and good separation
been used for separation of hydrogen isotope in zeolites, covalent performance of stable MOFs are still rare.
organic frameworks (COFs) and porous carbon materials due Not only the separation performance, but also structural
to the difference of de Broglie wavelength in hydrogen isotopes stability of MOFs plays a vital role in practical applications.
that creates different energy barrier at the entrance of the pores Among various MOF structures, some of these are very stable,
[1, 11–16]. In the confined system, ultra-low temperature can such as UiO [43, 44], MIL [45, 46] and ZIFs [47–49], which
lead to relatively high diffusion coefficient of the heavy isotope are widely used because of their stability. In 2014, Yaghi first
inside the porous materials, and resulting in the separation reported a series of ZIFs based on the chabazite (CHA) zeolite
of the hydrogen isotope [17–26]. Compared with the traditional topology using mixed-link, which exhibit high stability [47].
inorganic porous materials (zeolite, activated carbon, etc.), metal– Herein, we reported a new Cu-based MOF FJI-Y11 with CHA
organic frameworks (MOFs), also known as porous materials, topology by single ligand, which also exhibits excellent chemical
which were constructed with organic ligands and metal nodes or and thermal stability. Surprisingly, experimental and theoretical
clusters, have been extensively studied owing to their structural results show that the exposed open Cu sites on the paddlewheel

Address correspondence to Daqiang Yuan, ydq@fjirsm.ac.cn; Xiang He, hxiang@shu.edu.cn


Nano Res. 2021, 14(2): 518–525 519

clusters are not the priority adsorption sites. Indeed, the oxygen Ne (10/10/80, vol.%) or H2/D2/Ne (1/1/98, vol.%) gas mixtures,
atoms in the carboxyl group and the oxygen atoms connected dry FJI-Y11 powders (0.33 g) were packed into a stainless-steel
with methylene on the ligand play a major role in the adsorption column (diameter = 2 mm, length = 110 mm) with silica wool
of hydrogen and deuterium. Furthermore, a real-time cryogenic filling the void space. The column filled with the sample was
breakthrough device is used to verify the feasibility of FJI-Y11 degassed under dynamic vacuum at 373 K for 5 h and then a
for hydrogen isotope separation in a dynamic condition close constant Ne flow (10 mL/min) was used to purge the adsorption
to those used in the industrial environment. column for 2 h at 373 K to make certain that the adsorption
column was full of Ne. In order to ensure the accuracy of
2 Experimental the breakthrough experiments, the adsorption column and
pre-cooling line must be immersed into a liquid nitrogen tank
2.1 Materials for 20 min before tests. The raw gas mixtures at a flow rate of
10 mL/min for H2/D2/Ne (10/10/80, vol.%) or 15 mL/min for
The ligand 5,5'-[(1,4-phenylenebis(methylene))bis(oxy)]diiso- H2/D2/Ne (1/1/98, vol.%) were passed through the packed
phthalic acid (H4L) was prepared according to previously column. The gas flows were controlled by a mass flow meter
reported literature [50]. All other chemicals were analytical at the inlet and the composition of the effluent gas from the
reagent grade and used as purchased without further purification. adsorption column was continuously monitored by mass spec-
trometry. All of the samples for breakthrough experiments
2.2 Instruments
were regenerated with a Neon flow (10 mL/min) at 373 K for
The single-crystal X-ray diffraction data of FJI-Y11 was collected 30 min.
on a Bruker APEX-II diffractometer equipped with Mo Kα
radiation (λ = 0.71073 Å). Powder X-ray diffraction (PXRD) 2.6 Computational method
patterns were recorded on a Rigaku Mini Flex 600 X-ray The grand canonical Monte Carlo (GCMC) simulations were
diffractometer with Cu Kα radiation (λ = 1.5418 Å) in the performed using RASPA2.0 code [51] to explore H2 physisorption
scanning angle (2θ) range of 4°–50° at 293 K. Thermogravimetric in FJI-Y11. Lennard-Jones (LJ) parameters for the framework
analyses (TGA) were completed on a Netzsch STA 449C thermal atoms were taken from the DREIDING Force Field [52] except
analyzer at a heating rate of 10 °C/min in nitrogen. Elemental Cu atom from UFF Force Field [53], and the H2 LJ parameters
analyses (EA) for C, H and N were performed on a Vario EL from the TraPPE force field. Lorentz-Berthelot mixing rules
III analyzer. 1H NMR spectra were carried out on a Burker were used for all cross terms, and LJ interactions beyond 12.8 Å
AVANCE 400 spectrometer. were neglected. Feynman-Hibbs corrections were used in the
2.3 Synthesis of [Cu2(L)(H2O)2]n·nDMF·2nH2O (FJI-Y11) simulations. Coulomb interactions were calculated using partial
charges on the framework atoms. The Ewald sum method was
A mixture of Cu(NO3)2·3H2O (0.15 mmol), H4L (0.05 mmol), used to compute the electrostatic interactions. 100000 Monte
N,N-dimethylformamide (DMF) (2 mL), H2O (6 mL) and 2 M Carlo equilibration cycles were performed plus 100000 production
HNO3 (0.5 mL) was sealed in a 20 mL Teflon-lined reactor. cycles to calculate the ensemble averages. In one cycle, an
The mixture was ultrasonicated until homogeneous and average of N moves was performed, where N is the number of
then heated at 85 oC for 72 h. Green block crystals were molecules in the system. Monte Carlo moves used with equal
separated and washed with water (3 × 10 mL). The yield probability were translation, rotation, insertion, deletion, and
based on ligand is about 40%. Elemental microanalysis random reinsertion of an existing molecule at a new position.
for Cu2(C24H14O8)(H2O)2·C3H7NO·2H2O (C27H29NCu2O15), calc. The adsorption energy of H2 in MOF was investigated within
(%): C, 44.15; H, 3.98; N, 1.91; found (%): C, 44.08; H, 4.11; N, the framework of the density function theory (DFT) method.
1.90. IR (KBr pellet, cm−1): 3,242 w, 2,959 w, 2,918 w, 2,776 w, Due to the large number of atoms, the calculations were
1,659 s, 1,609 s, 1,550 s, 1,448 m, 1,374 s, 1,261 s, 1,132 m, performed based upon the primitive cells. Firstly, a reasonable
1,070 w, 1,020 w, 927 w, 786 m, 712 m. crystal structure was obtained by geometry optimization with
fixed cell parameters. Then, according to the GCMC results,
2.4 Gas sorption measurements
a hydrogen molecule was placed at the position where the
The adsorption isotherms of N2, H2 and D2 were measured adsorption was largest. Thus, two possible adsorbate-MOF
using an automatic volumetric adsorption apparatus (Micro- models (i.e. site I and site II models) were constructed for our
meritics ASAP 2020+) at 77 and 87 K. Before the gas sorption calculation. The PBE gradient-corrected functional with the
measurements, the sample was degassed under dynamic Grimme correction as well as the all-electronic DNP basis
vacuum (≤ 10 μmHg) at 100 °C for 10 h after solvent exchange sets was employed for all calculations. As for the geometry
with methanol and dichloromethane six times over three days, optimizations of the adsorbate-MOF models, the lattice
respectively. The specific surface area of FJI-Y11 was determined parameters and the atomic fraction positions of the MOF
using the Brunauer–Emmett–Teller (BET) model from the N2 crystal were kept fixed and only the H2 molecule was allowed
adsorption isotherm data between the relative pressure of 0.05 to relax. All of calculations were performed by the CASTEP and
and 0.3 at 77 K. Dmol3 modules integrated in the Material Studio 7.0 program
package. The adsorption energy was calculated in terms of the
2.5 Dynamic column breakthrough separation equation
experiments
Eads  EMOFH2  EMOF  EH2
The breakthrough curves were performed on a real-time low-
temperature apparatus coupled with a mass spectrometer where EMOFH stands for the energy of the optimized adsorbate-
2
(PFEIFFER VACUUM) using the gas mixtures of H2/D2/Ne MOF structure, and EMOF, and EH denote the energies of the
2
(10/10/80, vol.%) and H2/D2/Ne (1/1/98, vol.%) at 77 K and bare MOF structure and the isolated H2 molecule, respectively.
101 kPa (Fig. S1 in the Electronic Supplementary Material According to this equation, the more negative adsorption
(ESM)). In a typical breakthrough experiment for the H2/D2/ energy means more favorable binding.

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research


520 Nano Res. 2021, 14(2): 518–525

3 Results and discussion seen that the channels of the framework are not regularly
distributed along the axial direction. For activated FJI-Y11, its
3.1 Structural analysis of FJI-Y11 global cavity diameter and largest cavity diameter have the same
value of 8.4 Å, and the pore limiting diameter is 3.8 Å [54, 55].
X-ray crystallography shows that FJI-Y11 crystallizes in the To our knowledge, FJI-Y11 is the first case of MOF with CHA
trigonal space group R 3 . The asymmetric unit consists of two topology based on single ligand. The guest solvent molecules,
crystallographically independent Cu(II) ions, one completely including DMF and H2O, are highly disordered in the channels,
deprotonated organic ligand and two coordinated water. As so the diffraction intensity contributed by the disorder solvents
shown in Fig. 1(b), each Cu(II) center adopts a square pyramidal molecules has been squeezed with Platon [56]. The reasonable
geometry, and the equatorial Cu−O distances span the range of solvent composition was calculated on the basis of the
1.922(7)–1.990(7) Å, which suggests strong metal−carboxylate thermogravimetric weight loss analysis and elemental analysis.
bonding. The axially bound water molecules show the Cu1−O12
and Cu2−O11 distances to be 2.161(7) and 2.150(7) Å, respec- 3.2 Stability
tively. Two Cu(II) ions are linked by four bridging carboxylates
In order to evaluate the thermal stability of FJI-Y11, TGA were
to form the well-known [Cu2(CO2)4(H2O)2] paddlewheel unit,
carried out under an N2 environment (Fig. S3 in the ESM).
with the Cu···Cu separation of 2.6277 Å. Each L is coordinated
The result shows that FJI-Y11 can be stable up to about 275 °C
to four paddlewheel units, which shows μ8-η1:η1:η1:η1:η1:η1:η1:η1
and has nearly 19% weight loss before 275 °C, corresponding
coordination mode. The dihedral angle between isophthalic
to the removal of one DMF molecule and four water molecules.
rings is 84.97°. The central phenyl ring twists away from
The thermal stability of FJI-Y11 can also be verified by variable
the two isophthalic rings with dihedral angles are 81.88° and
temperature powder X-ray diffraction (VT-PXRD) results
81.24°, respectively. So the Cu2(CO2)4(H2O)2 unit can be viewed
(Fig. 2(a)). The slight difference in collapse temperature between
as a 4-connected node. While each L coordinates to four
TGA and VT-PXRD maybe because the VT-PXRD pattern is
paddlewheel secondary building units (SBUs) can also act as
obtained in air while the TGA spectrum is measured in N2.
a 4-connected node. Extension of the coordination geometry
It should be noted that FJI-Y11 not only exhibits good water
around these 4-connected nodes will produce the final CHA
stability but also has acid/alkali stability. When treated by
framework with three-dimensional percolated channels
boiling water for one week or acid/alkali solution (pH = 2–12)
(Figs. 1(c) and 1(d)). Figure S2 in the ESM shows the pore
for three days at 100 °C, FJI-Y11 still retains its stability of the
surface structure of activated FJI-Y11, and it can be clearly
framework, which can be proved by the PXRD results (Fig. 2(b)
and Fig. S5 in the ESM). The excellent stable behaviour of FJI-Y11
is crucial for its further practical applications. Generally speaking,
Cu-based MOFs only with traditional paddle-wheel structure
is usually hydrolysed in the presence of water, and only a few
known MOFs have such excellent chemical stability [57]. Even
for the famous material HKUST-1, its crystallinity was lost after
treated with boiling aqueous solution only one hour (Fig. S14
in the ESM). Note that the stability of FJI-Y11 and HKUST-1
discussed above is for the synthetic framework where water
molecules are coordinated to the Cu center in the paddlewheel
cluster. The coordinated water molecules can be removed during
activation, resulting in the color change from the initial green
to the final dark blue due to the change of crystal field (Fig. S7
in the ESM). We performed the same acid/alkali tests on the
activated FJI-Y11 as the original hydrated form, and the results
showed that the activated FJI-Y11 still maintained its stability
(Fig. S6 in the ESM). We speculate this unusual chemical stability
of FJI-Y11 is due to its small pore size. On rehydration, the
Figure 1 Structure introduction of FJI-Y11. (a) The ligand H4L and SBU pore limiting diameter is narrowed from 3.8 to 3.3 Å as each
for constructing FJI-Y11. (b) The coordination environments of Cu(II) open Cu site is re-occupied by a water molecule (Fig. S8 in the
ion as the paddlewheel dimer. (c) The three-dimensional channels. (d) ESM), which may prevent more water molecules from further
CHA topology of FJI-Y11. attacking metal–carboxylate oxygen bonds, resulting in high

Figure 2 Stability of FJI-Y11. (a) VT-PXRD patterns measured in air; (b) PXRD patterns after treatment with acid/alkali solution (pH = 2–12) for 3 days
at 100 °C.

| www.editorialmanager.com/nare/default.asp
Nano Res. 2021, 14(2): 518–525 521

chemical stability of FJI-Y11.

3.3 Permanent porosity and N2 adsorption


The crystallinity of FJI-Y11 was retained after the removal of
guest molecules and coordinated water molecules on the Cu
paddlewheel cluster, as indicated by PXRD result (Fig. S4 in
the ESM). The Zeo++ calculation (probe radius = 1.86 Å) reveals
that the potential solvent-accessible void fraction of activated
FJI-Y11 is 36.3%, which encourages us to investigate its permanent
porosity. As shown in Fig. 3, the N2 adsorption curve exhibits
typical type I adsorption isotherm with a maximum uptake of
190 cm3/g at 1 bar. The N2 adsorption curve increases steeply
below relative pressure 0 and saturates at relative pressure 0.2.
In the desorption process, the N2 is desorbed totally when
the pressure reduces, which clearly indicates the micropores
characteristic of FJI-Y11. The BET and Langmuir surface
areas are 719 and 828 m2/g, respectively. Accordingly, the total Figure 4 H2 and D2 sorption isotherms of FJI-Y11 at 77 and 87 K (filled
pore volume computed by the nitrogen isotherm is 0.30 cm3/g symbols: adsorption; open symbols: desorption).
at p/p0 = 0.99. Based on the N2 adsorption data, pore size
distribution calculated by the model of nonlocal density functional isotope separation, because the adsorption capacity is also an
theory (NLDFT) shows a narrow distribution of micropores important parameter in practical applications. In order to
(about 5.9 Å), which is closed to the value calculated by Zeo++. evaluate the affinity between framework and hydrogen isotopes,
the adsorption enthalpies of hydrogen and deuterium were
3.4 Adsorption and separation properties of hydrogen calculated by Viral equation based on the adsorption isotherm
isotope data at 77 and 87 K with values of 7.13 and 7.88 kJ/mol at zero
The permanent micropores of FJI-Y11 prompted us to study coverage, respectively (Fig. 5(a)). It is worth noting that hydrogen
its gas uptake capacity, especially for hydrogen and deuterium. isotopes adsorption capacities of HKUST-1 are much larger
Moreover, we investigated hydrogen and deuterium adsorption than that of FJI-Y11 (Fig. S15 in the ESM), but FJI-Y11 possesses
of FJI-Y11 at 77 and 87 K, respectively. As shown in Fig. 4, higher adsorption enthalpies, which indicates that FJI-Y11 has
activated FJI-Y11 show remarkable hydrogen and deuterium stronger interaction with hydrogen isotopes (Fig. S16 in the
uptakes, the adsorption capacity up to 183 and 205 cm3/g at ESM) [64]. Furthermore, the ideal adsorption solution theory
77 K and 1 bar, and the values reaching 146 and 167 cm3/g at (IAST) was used to calculate the selectivity of equimolar D2/H2
87 K and 1 bar. The hydrogen uptake exceeds many famous mixtures at atmospheric pressure. As shown in Fig. 5(b), the
selectivity of FJI-Y11 is 1.76 at 77 K and 100 kPa, while the
MOFs at low pressure, such as MOF-74-Zn (144 cm3/g) [58],
MIL-53 (Al) (21 cm3/g) [59], UiO-66 (151 cm3/g) and ZIF-8 selectivity of HKUST-1 is only 1.42 under the same conditions
(143 cm3/g) [60–63]. Compared with similar BET and pore (Fig. S17 in the ESM) [65, 66]. Compared with HKUST-1, the
better selective property of FJI-Y11 may be caused by its relatively
size of pores materials, the hydrogen adsorption amount of
smaller pore size. Meanwhile, FJI-Y11 still shows better selective
FJI-Y11 is at a high level. The high hydrogen and deuterium
uptakes of FJI-Y11 may be ascribed to the adsorption sites and performance than that of zeolite 5A which has similar pore
size (Figs. S19 and S20 in the ESM), indicating that the adsorption
suitable pore space. The adsorption capacity of deuterium gas
sites on FJI-Y11 play an essential role in the separation of
is higher than that of hydrogen, which is owing to the fact that
the heavier isotope has lower zero energy and smaller de Broglie hydrogen isotope. Thus, the highly selective adsorption of D2
over H2 for FJI-Y11 further illustrates that the combination of
wavelength as compared to hydrogen. Additionally, high
suitable pore space and strong binding sites has a positive effect
hydrogen and deuterium uptakes are beneficial for hydrogen
on the selective adsorption of D2. Therefore, FJI-Y11 could be
acted as a candidate material for hydrogen isotope separation.
It is noted that the reported hydrogen adsorption performance
of HKUST-1 varies quite substantially, for example, Bordiga et
al. [67] and Li et al. [68] both reported hydrogen adsorption
capacity of 6 mmol/g, while Yaghi et al. reported hydrogen uptake
of 7.9 mmol/g in one paper and 12.4 mmol/g in another paper
[69–71]. All the results about HKUST-1 in this paper are based
on the highly crystalline material with hydrogen uptake of
12.4 mmol/g.

3.5 The hydrogen isotopes adsorption sites in activated


FJI-Y11
To confirm hydrogen and deuterium adsorption sites, the
hydrogen isotopes adsorption behaviours of activated FJI-Y11
was studied by multilevel computational simulations. The GCMC
simulations show that the simulated hydrogen and deuterium
adsorption isotherms of activated FJI-Y11 are in agreement
Figure 3 Nitrogen sorption isotherm of FJI-Y11 at 77 K (filled symbols: with the experimental data at 77 K (Figs. S9 and S10 in the
adsorption; open symbols: desorption); the inset graph shows its pore size ESM), whereas the adsorption capacities are almost the same
distribution analysed by NLDFT method. at atmospheric pressure. As shown in the Figs. S11 and S12 in

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research


522 Nano Res. 2021, 14(2): 518–525

Figure 5 (a) The isosteric heat of D2 (black) and H2 (red) for FJI-Y11. (b) IAST selectivity of FJI-Y11 for 1:1 D2/H2 mixture at 77 K (black) and 87 K (red).

the ESM, the concentration of deuterium is higher than that of activated FJI-Y11 with a total flow of 10 mL/min. As expected,
hydrogen at the same lower pressure, which indicates that the challenging separation of hydrogen isotope was realized by
deuterium is easier to enter and adsorb in the hole. Careful using FJI-Y11. From the breakthrough curve, hydrogen was first
analysis of the simulated density picture of adsorbed hydrogen to eluted through the packed column, whereas the adsorption
molecules shows that there are two preferential adsorption column retained deuterium for a long time of 17 min/g before
sites (Figs. 6(b) and 6(c)). The DFT method is used to further the breakthrough of D2 (Fig. 7(a)). The amount of D2 enriched
optimize the two adsorption sites. Abnormally, the first in the adsorption bed from the mixed gas of H2/D2/Ne
adsorption sites of simulation are located on the oxygen atoms (10/10/80, vol.%) was up to 4.1 mmol/g, and the separation factor
of Cu paddlewheel clusters rather than the open Cu sites. reached 1.4 which was counted based on the breakthrough
In site I (Fig. 6(b)), hydrogen molecule is closest to the oxygen experiments at 77 K and 1 bar. Besides, we also studied the
atom on the Cu paddlewheel cluster with H–O distance of H2/D2 separation performance using the mixture of H2/D2/Ne
2.53 Å, while the distance of Cu–H is 3.20 Å. As shown in (1/1/98, vol.%) with a flow rate of 15 mL/min and the retention
Fig. 6(c), the shortest H–O distance of 2.84 Å in site II is time of deuterium can be as high as 70 min/g (Fig. 7(b)). For
between hydrogen molecule and the oxygen atom connected comparison, breakthrough experiments for HKUST-1 and zeolite
with methylene on the ligand, which shows that the additional 5 A were also carried out. As shown in Figs. 7(c) and 7(d), the
O atoms on the ligands (not the O atoms of carboxyl group) retention time of deuterium was only 10 and 6 min/g for
have a positive effect on further promoting hydrogen adsorption. HKUST-1 and zeolite 5A, respectively. And the trend of the
The calculated hydrogen adsorption energies of adsorption retention times about three materials for the H2/D2/Ne (1/1/98,
site I and site II are 21.10 and 19.29 kJ/mol, respectively. vol.%) mixture is the same as the mixture of H2/D2/Ne
(10/10/80, vol.%) (Figs. S18 and S21 in the ESM). Compared with
3.6 Dynamic column breakthrough experiments HKUST-1 and zeolite 5A, the better separation performance
To further investigate the D2/H2 separation property of FJI-Y11 of H2/D2 for FJI-Y11 could be ascribed to the synergistic effect
in the practical application, dynamic breakthrough experiments of the strong adsorption sites and suitable pore space. It is worth
were performed at 77 K and 1 bar, in which a mixed gas of noting that the pore limiting diameter (3.8 Å) of FJI-Y11 is
H2/D2/Ne (10/10/80, vol.%) flowed over the packed column of small enough, but the KQS effect is difficult to be observed

Figure 6 (a) Slices through the calculated potential fields for adsorbed hydrogen in activated FJI-Y11 under 77 K. The slice is viewed along the
crystallographic c axis. The simulated adsorption (b) site I and (c) site II in activated FJI-Y11 obtained from the GCMC simulation and DFT optimization.
Some atoms are omitted for clarity.

| www.editorialmanager.com/nare/default.asp
Nano Res. 2021, 14(2): 518–525 523

Figure 7 The dynamic breakthrough curves of FJI-Y11 at 77 K for the mixed gases of (a) H2/D2/Ne (10/10/80 vol.%); (b) H2/D2/Ne (1/1/98 vol.%). The
dynamic breakthrough curve for a H2/D2/Ne (10/10/80 vol.%) mixture at 77 K in a packed column with (c) HKUST-1 and (d) zeolite 5A.

due to the relatively high temperature used for the breakthrough 4 Conclusions
experiment [32]. However, there is no doubt that the suitable
pore space does play a positive role in the process of hydrogen In summary, we have successfully synthesized a microporous
isotope separation, which has been confirmed by above Cu-based MOF which has rare chabazite topology and exhibits
experiments. excellent chemical and thermal stability. This MOF possesses
In practical application, adsorbent should have good structural high hydrogen and deuterium adsorption amounts due to
stability and regeneration ability [29]. Thus, breakthrough the presence of two special adsorption sites. Interestingly,
experiments of the ternary mixed gas of H2/D2/Ne (10/10/80 the adsorption site I is not the exposed open Cu site on the
vol.%) were executed to evaluate the recycling performance of paddlewheel cluster due to the mismatch between the direction
FJI-Y11. As shown in Fig. 8, there is no significant loss in the of open metal sites and the diffusion path of gas molecules.
D2 adsorption capability and breakthrough time for FJI-Y11. Furthermore, breakthrough experiments at high temperature
In addition, FJI-Y11 retained its crystallinity and its BET did not of 77 K certifies that the heavy hydrogen isotope can be effectively
change obviously after three cycling breakthrough experiments separated and stored inside the pores of FJI-Y11 by combining
(Figs. S4 and S13 in the ESM). In short, enriched deuterium can with suitable space and adsorption sites in one framework. As far
be recovered at a relatively high level of purity as compared to as we know, this is the first time that using MOF as adsorbent
the raw mixture during the regeneration process, and all these bed to explore hydrogen isotope separation by breakthrough
results suggest that FJI-Y11 is a promising candidate material experiment. Overall, high hydrogen isotopes adsorption capacity,
for H2/D2 purification. good gas separation property and excellent regenerative stability
enable that FJI-Y11 as excellent adsorbent for separating hydrogen
isotope may be used in the industrial application. In addition,
breakthrough separation technology as the promising gas
separation technology with the advantage of low-cost and
low-energy can be further extended to solve the challenge of
separating hydrogen isotope. By introducing strong binding
sites into the micropores materials and controlling the pore
size at the same time, highly selective adsorption of D2 over H2
will be possible. The rational design of material FJI-Y11 will
provide a new synthetic strategy for the future hydrogen isotope
separation materials.

Acknowledgements
This work was financially supported by the Strategic Priority
Research Program of CAS (No. XDB20000000), the Key Research
Figure 8 The recyclability of FJI-Y11 in multiple mixed-gas column Program of Frontier Sciences, CAS (No. QYZDB-SSW-SLH019),
breakthrough tests for a H2/D2/Ne (10/10/80 vol.%) mixture with a total and the Natural Nature Science Foundation of China (Nos.
flow of 10 mL/min at 77 K. 21771177, 51603206 and 21203117).

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research


524 Nano Res. 2021, 14(2): 518–525

Electronic Supplementary Material: Supplementary material Hydrogen Energy 2014, 39, 4437–4446.
(attached with all the supporting figures mentioned in this [19] Oh, H.; Kalidindi, S. B.; Um, Y.; Bureekaew, S.; Schmid, R.; Fischer,
work) is available in the online version of this article at https:// R. A.; Hirscher, M. A cryogenically flexible covalent organic
doi.org/10.1007/s12274-019-2571-9. framework for efficient hydrogen isotope separation by quantum
sieving. Angew. Chem., Int. Ed. 2013, 52, 13219–13222.
[20] Krkljus, I.; Steriotis, T.; Charalambopoulou, G.; Gotzias, A.; Hirscher,
References M. H2/D2 adsorption and desorption studies on carbon molecular
sieves with different pore structures. Carbon 2013, 57, 239–247.
[1] Beenakker, J. J. M.; Borman, V. D.; Krylov, S. Y. Molecular transport
[21] Cai, J. J.; Xing, Y. L.; Zhao, X. B. Quantum sieving: Feasibility and
in subnanometer pores: Zero-point energy, reduced dimensionality
challenges for the separation of hydrogen isotopes in nanoporous
and quantum sieving. Chem. Phys. Lett. 1995, 232, 379–382.
materials. RSC Adv. 2012, 2, 8579–8586.
[2] Basmadjian, D. Adsorption equilibria of hydrogen, deuterium, and
[22] Nguyen, T. X.; Jobic, H.; Bhatia, S. K. Microscopic observation of
their mixtures. Part I. Can. J. Chem. 1960, 38, 141–148.
kinetic molecular sieving of hydrogen isotopes in a nanoporous
[3] Johnston, H. L.; Long, E. A. Heat capacity curves of the simpler
material. Phys. Rev. Lett. 2010, 105, 085901.
gases. vi. Rotational heat capacity curves of molecular deuterium and
[23] Kotoh, K.; Takashima, S.; Sakamoto, T.; Tsuge, T. Multi-component
of deuterium hydride. The equilibrium between the ortho and para
behaviors of hydrogen isotopes adsorbed on synthetic zeolites 4A
forms of deuterium. Free energy, total energy, entropy, heat capacity
and 5A at 77.4 K and 87.3 K. Fusion Eng. Des. 2010, 85, 1928–1934.
and dissociation of H2H2 and of H1H2, to 3000 °K. J. Chem. Phys.
[24] Garberoglio, G. Quantum sieving in organic frameworks. Chem. Phys.
1934, 2, 389–395.
Lett. 2009, 467, 270–275.
[4] Yaris, R.; Sams, J. R. Jr. Quantum treatment of the physical adsorption
[25] Panella, B.; Hirscher, M.; Ludescher, B. Low-temperature thermal-
of isotopic species. J. Chem. Phys. 1962, 37, 571–576.
desorption mass spectroscopy applied to investigate the hydrogen
[5] Povinec, P. P.; Bokuniewicz, H.; Burnett, W. C.; Cable, J.; Charette,
adsorption on porous materials. Micropor. Mesopor. Mater. 2007,
M.; Comanducci, J. F.; Kontar, E. A.; Moore, W. S.; Oberdorfer, J. 103, 230–234.
A.; de Oliveira, J. et al. Isotope tracing of submarine groundwater [26] Zhao, X. B.; Villar-Rodil, S.; Fletcher, A. J.; Thomas, K. M. Kinetic
discharge offshore Ubatuba, Brazil: Results of the IAEA-UNESCO isotope effect for H2 and D2 quantum molecular sieving in adsorption/
SGD project. J. Environ. Radioact. 2008, 99, 1596–1610. desorption on porous carbon materials. J. Phys. Chem. B 2006, 110,
[6] Keppler, F.; Hamilton, J. T. G.; McRoberts, W. C.; Vigano, I.; Braß, 9947–9955.
M.; Röckmann, T. Methoxyl groups of plant pectin as a precursor of [27] Zhou, H. C.; Long, J. R.; Yaghi, O. M. Introduction to metal-organic
atmospheric methane: Evidence from deuterium labelling studies. frameworks. Chem. Rev. 2012, 112, 673–674.
New Phytol. 2008, 178, 808–814. [28] Suh, M. P.; Park, H. J.; Prasad, T. K.; Lim, D. W. Hydrogen storage
[7] Zaccai, G. How soft is a protein? A protein dynamics force constant in metal-organic frameworks. Chem. Rev. 2012, 112, 782–835.
measured by neutron scattering. Science 2000, 288, 1604–1607. [29] Kapelewski, M. T.; Runčevski, T.; Tarver, J. D.; Jiang, H. Z. H.;
[8] Büldt, G.; Gally, H. U.; Seelig, A.; Seelig, J.; Zaccai, G. Neutron Hurst, K. E.; Parilla, P. A.; Ayala, A.; Gennett, T.; FitzGerald, S. A.;
diffraction studies on selectively deuterated phospholipid bilayers. Brown, C. M. et al. Record high hydrogen storage capacity in the
Nature 1978, 271, 182–184. metal–organic framework Ni2(m-dobdc) at near-ambient temperatures.
[9] Shi, C. W.; Fricke, P.; Lin, L.; Chevelkov, V.; Wegstroth, M.; Giller, K.; Chem. Mater. 2018, 30, 8179–8189.
Becker, S.; Thanbichler, M.; Lange, A. Atomic-resolution structure [30] Noguchi, D.; Tanaka, H.; Kondo, A.; Kajiro, H.; Noguchi, H.;
of cytoskeletal bactofilin by solid-state NMR. Sci. Adv. 2015, 1, Ohba, T.; Kanoh, H.; Kaneko, K. Quantum sieving effect of three-
e1501087. dimensional Cu-based organic framework for H2 and D2. J. Am.
[10] Tanaka, H.; Noguchi, D.; Yuzawa, A.; Kodaira, T.; Kanoh, H.; Kaneko, Chem. Soc. 2008, 130, 6367–6372.
K. Quantum effects on hydrogen isotopes adsorption in nanopores. [31] Teufel, J.; Oh, H.; Hirscher, M.; Wahiduzzaman, M.; Zhechkov, L.;
J. Low Temp. Phys. 2009, 157, 352–373. Kuc, A.; Heine, T.; Denysenko, D.; Volkmer, D. MFU-4—A metal-
[11] Salazar, J. M.; Lectez, S.; Gauvin, C.; Macaud, M.; Bellat, J. P.; organic framework for highly effective H2/D2 separation. Adv. Mater.
Weber, G.; Bezverkhyy, I.; Simon, J. M. Adsorption of hydrogen 2013, 25, 635–639.
isotopes in the zeolite NaX: Experiments and simulations. Int. J. [32] Paschke, B.; Denysenko, D.; Bredenkötter, B.; Sastre, G.; Wixforth,
Hydrogen Energy 2017, 42, 13099–13110. A.; Volkmer, D. Dynamic studies on kinetic H2/D2 quantum sieving
[12] Physick, A. J. W.; Wales, D. J.; Owens, S. H. R.; Shang, J.; Webley, in a narrow pore metal-organic framework grown on a sensor chip.
P. A.; Mays, T. J.; Ting, V. P. Novel low energy hydrogen–deuterium Chem.—Eur. J. 2019, 25, 10803–10807.
isotope breakthrough separation using a trapdoor zeolite. Chem. Eng. [33] Weinrauch, I.; Savchenko, I.; Denysenko, D.; Souliou, S. M.; Kim,
J. 2016, 288, 161–168. H. H.; Le Tacon, M.; Daemen, L. L.; Cheng, Y.; Mavrandonakis, A.;
[13] Zhao, X. B.; Xiao, B.; Fletcher, A. J.; Thomas, K. M. Hydrogen Ramirez-Cuesta, A. J. et al. Capture of heavy hydrogen isotopes in a
adsorption on functionalized nanoporous activated carbons. J. Phys. metal-organic framework with active Cu(I) sites. Nat. Commun.
Chem. B 2005, 109, 8880–8888. 2017, 8, 14496.
[14] Kumar, A. V. A.; Bhatia, S. K. Quantum effect induced reverse kinetic [34] Chen, B. L.; Zhao, X. B.; Putkham, A.; Hong, K. L.; Lobkovsky, E. B.;
molecular sieving in microporous materials. Phys. Rev. Lett. 2005, Hurtado, E. J.; Fletcher, A. J.; Thomas, K. M. Surface interactions
95, 245901. and quantum kinetic molecular sieving for H2 and D2 adsorption on
[15] Tanaka, H.; Kanoh, H.; Yudasaka, M.; Iijima, S.; Kaneko, K. Quantum a mixed metal-organic framework material. J. Am. Chem. Soc. 2008,
effects on hydrogen isotope adsorption on single-wall carbon 130, 6411–6423.
nanohorns. J. Am. Chem. Soc. 2005, 127, 7511–7516. [35] Kim, J. Y.; Balderas-Xicohténcatl, R.; Zhang, L. D.; Kang, S. G.;
[16] Stéphanie-Victoire, F.; Goulay, A. M.; de Lara, E. C. Adsorption and Hirscher, M.; Oh, H.; Moon, H. R. Exploiting diffusion barrier and
coadsorption of molecular hydrogen isotopes in zeolites. 1. Isotherms chemical affinity of metal-organic frameworks for efficient hydrogen
of H2, HD, and D2 in NaA by thermomicrogravimetry. Langmuir isotope separation. J. Am. Chem. Soc. 2017, 139, 15135–15141.
1998, 14, 7255–7259. [36] FitzGerald, S. A.; Pierce, C. J.; Rowsell, J. L. C.; Bloch, E. D.;
[17] Cendagorta, J. R.; Powers, A.; Hele, T. J. H.; Marsalek, O.; Bačić, Z.; Mason, J. A. Highly selective quantum sieving of D2 from H2 by a
Tuckerman, M. E. Competing quantum effects in the free energy metal-organic framework as determined by gas manometry and
profiles and diffusion rates of hydrogen and deuterium molecules infrared spectroscopy. J. Am. Chem. Soc. 2013, 135, 9458–9464.
through clathrate hydrates. Phys. Chem. Chem. Phys. 2016, 18, [37] FitzGerald, S. A.; Burkholder, B.; Friedman, M.; Hopkins, J. B.;
32169–32177. Pierce, C. J.; Schloss, J. M.; Thompson, B.; Rowsell, J. L. C. Metal-
[18] Chu, X. Z.; Cheng, Z. P.; Xiang, X. X.; Xu, J. M.; Zhao, Y. J.; Zhang, specific interactions of H2 adsorbed within isostructural metal-organic
W. G.; Lv, J. S.; Zhou, Y. P.; Zhou, L.; Moon, D. K. et al. Separation frameworks. J. Am. Chem. Soc. 2011, 133, 20310–20318.
dynamics of hydrogen isotope gas in mesoporous and microporous [38] FitzGerald, S. A.; Hopkins, J.; Burkholder, B.; Friedman, M.; Rowsell,
adsorbent beds at 77 K: SBA-15 and zeolites 5A, Y, 10X. Int. J. J. L. C. Quantum dynamics of adsorbed normal- and para-H2, HD,

| www.editorialmanager.com/nare/default.asp
Nano Res. 2021, 14(2): 518–525 525

and D2 in the microporous framework MOF-74 analyzed using identifying geometrically diverse sets of crystalline porous materials.
infrared spectroscopy. Phys. Rev. B 2010, 81, 104305. J. Chem. Inf. Model. 2012, 52, 308–318.
[39] Oh, H.; Hirscher, M. Quantum sieving for separation of hydrogen [56] Spek, A. L. PLATON SQUEEZE: A tool for the calculation of the
isotopes using MOFs. Eur. J. Inorg. Chem. 2016, 2016, 4278–4289. disordered solvent contribution to the calculated structure factors.
[40] Oh, H.; Park, K. S.; Kalidindi, S. B.; Fischer, R. A.; Hirscher, M. Acta Crystallogr. C Struct. Chem. 2015, 71, 9–18.
Quantum cryo-sieving for hydrogen isotope separation in microporous [57] McHugh, L. N.; McPherson, M. J.; McCormick, L. J.; Morris, S. A.;
frameworks: An experimental study on the correlation between Wheatley, P. S.; Teat, S. J.; McKay, D.; Dawson, D. M.; Sansome, C.
effective quantum sieving and pore size. J. Mater. Chem. A 2013, 1, E. F.; Ashbrook, S. E. et al. Hydrolytic stability in hemilabile
3244–3248. metal-organic frameworks. Nat. Chem. 2018, 10, 1096–1102.
[41] Liu, M.; Zhang, L. D.; Little, M. A.; Kapil, V.; Ceriotti, M.; Yang, S. [58] Gygi, D.; Bloch, E. D.; Mason, J. A.; Hudson, M. R.; Gonzalez, M.
Y.; Ding, L. F.; Holden, D. L.; Balderas-Xicohténcatl, R.; He, D. L. I.; Siegelman, R. L.; Darwish, T. A.; Queen, W. L.; Brown, C. M.;
et al. Barely porous organic cages for hydrogen isotope separation. Long, J. R. Hydrogen storage in the expanded pore metal–organic
Science 2019, 366, 613–620. frameworks M2(dobpdc) (M = Mg, Mn, Fe, Co, Ni, Zn). Chem. Mater.
[42] Oh, H.; Savchenko, I.; Mavrandonakis, A.; Heine, T.; Hirscher, M. 2016, 28, 1128–1138.
Highly effective hydrogen isotope separation in nanoporous metal- [59] Kim, J. Y.; Zhang, L. D.; Balderas-Xicohténcatl, R.; Park, J.; Hirscher,
organic frameworks with open metal sites: Direct measurement and M.; Moon, H. R.; Oh, H. Selective hydrogen isotope separation via
theoretical analysis. ACS Nano 2014, 8, 761–770. breathing transition in MIL-53(Al). J. Am. Chem. Soc. 2017, 139,
[43] Wang, H.; Dong, X. L.; Lin, J. Z.; Teat, S. J.; Jensen, S.; Cure, J.; 17743–17746.
Alexandrov, E. V.; Xia, Q. B.; Tan, K.; Wang, Q. N. et al. [60] Zhao, Q.; Yuan, W.; Liang, J. M.; Li, J. P. Synthesis and hydrogen
Topologically guided tuning of Zr-MOF pore structures for highly storage studies of metal−organic framework UiO-66. Int. J.
selective separation of C6 alkane isomers. Nat. Commun. 2018, 9, Hydrogen Energy 2013, 38, 13104–13109.
1745. [61] Ren, J. W.; Langmi, H. W.; North, B. C.; Mathe, M.; Bessarabov, D.
[44] Hu, Z. G.; Peng, Y. W.; Kang, Z. X.; Qian, Y. H.; Zhao, D. A Modulated synthesis of zirconium-metal organic framework (Zr-MOF)
Modulated Hydrothermal (MHT) approach for the facile synthesis for hydrogen storage applications. Int. J. Hydrogen Energy 2014,
of UiO-66-Type MOFs. Inorg. Chem. 2015, 54, 4862–4868. 39, 890–895.
[45] Cooper, L.; Guillou, N.; Martineau, C.; Elkaim, E.; Taulelle, F.; [62] Chavan, S.; Vitillo, J. G.; Gianolio, D.; Zavorotynska, O.; Civalleri, B.;
Serre, C.; Devic, T. ZrIV coordination polymers based on a naturally Jakobsen, S.; Nilsen, M. H.; Valenzano, L.; Lamberti, C.; Lillerud, K.
occurring phenolic derivative. Eur. J. Inorg. Chem. 2014, 2014, P. et al. H2 storage in isostructural UiO-67 and UiO-66 MOFs. Phys.
6281–6289. Chem. Chem. Phys. 2012, 14, 1614–1626.
[46] Férey, G.; Mellot-Draznieks, C.; Serre, C.; Millange, F.; Dutour, J.; [63] Zhou, M.; Wang, Q.; Zhang, L.; Liu, Y. C.; Kang, Y. Adsorption sites
Surblé, S.; Margiolaki, I. A chromium terephthalate-based solid with of hydrogen in zeolitic imidazolate frameworks. J. Phys. Chem. B
unusually large pore volumes and surface area. Science 2005, 309, 2009, 113, 11049–11053.
2040–2042. [64] Chui, S. S. Y.; Lo, S. M. F.; Charmant, J. P. H.; Orpen, A. G.;
[47] Nguyen, N. T. T.; Furukawa, H.; Gándara, F.; Nguyen, H. T.; Williams, I. D. A chemically functionalizable nanoporous material
Cordova, K. E.; Yaghi, O. M. Selective capture of carbon dioxide under [Cu3(TMA)2(H2O)3]n. Science 1999, 283, 1148–1150.
humid conditions by hydrophobic chabazite-type zeolitic imidazolate [65] Krkljus, I.; Hirscher, M. Characterization of hydrogen/deuterium
frameworks. Angew. Chem., Int. Ed. 2014, 53, 10645–10648. adsorption sites in nanoporous Cu–BTC by low-temperature thermal-
[48] Hayashi, H.; Côté, A. P.; Furukawa, H.; O'Keeffe, M.; Yaghi, O. M. desorption mass spectroscopy. Micropor. Mesopor. Mater. 2011,
Zeolite a imidazolate frameworks. Nat. Mater. 2007, 6, 501–506. 142, 725–729.
[49] Park, K. S.; Ni, Z.; Côté, A. P.; Choi, J. Y.; Huang, R. D.; Uribe-Romo, [66] Peterson, V. K.; Brown, C. M.; Liu, Y.; Kepert, C. J. Structural study
F. J.; Chae, H. K.; O'Keeffe, M.; Yaghi, O. M. Exceptional chemical of D2 within the trimodal pore system of a metal organic framework.
and thermal stability of zeolitic imidazolate frameworks. Proc. Natl. J. Phys. Chem. C 2011, 115, 8851–8857.
Acad. Sci. USA 2006, 103, 10186–10191. [67] Prestipino, C.; Regli, L.; Vitillo, J. G.; Bonino, F.; Damin, A.;
[50] Pan, Y. J.; Ford, W. T. Dendrimers with alternating amine and ether Lamberti, C.; Zecchina, A.; Solari, P. L.; Kongshaug, K. O.; Bordiga,
generations. J. Org. Chem. 1999, 64, 8588–8593. S. Local structure of framework Cu(II) in HKUST-1 metallorganic
[51] Dubbeldam, D.; Calero, S.; Ellis, D. E.; Snurr, R. Q. RASPA: Molecular framework: Spectroscopic characterization upon activation and
simulation software for adsorption and diffusion in flexible nanoporous interaction with adsorbates. Chem. Mater. 2006, 18, 1337–1346.
materials. Mol. Simulat. 2016, 42, 81–101. [68] Lee, J.; Li, J.; Jagiello, J. Gas sorption properties of microporous
[52] Mayo, S. L.; Olafson, B. D.; Goddard, W. A. DREIDING: A generic metal organic frameworks. J. Solid State Chem. 2005, 178, 2527–2532.
force field for molecular simulations. J. Phys. Chem. 1990, 94, [69] Rowsell, J. L. C.; Yaghi, O. M. Effects of functionalization, catenation,
8897–8909. and variation of the metal oxide and organic linking units on
[53] Rappe, A. K.; Casewit, C. J.; Colwell, K. S.; Goddard III, W. A.; the low-pressure hydrogen adsorption properties of metal-organic
Skiff, W. M. UFF, a full periodic table force field for molecular frameworks. J. Am. Chem. Soc. 2006, 128, 1304–1315.
mechanics and molecular dynamics simulations. J. Am. Chem. Soc. [70] Wong-Foy, A. G.; Matzger, A. J.; Yaghi, O. M. Exceptional H2 saturation
1992, 114, 10024–10035. uptake in microporous metal-organic frameworks. J. Am. Chem. Soc.
[54] Willems, T. F.; Rycroft, C. H.; Kazi, M.; Meza, J. C.; Haranczyk, M. 2006, 128, 3494–3495.
Algorithms and tools for high-throughput geometry-based analysis [71] Xiao, B.; Wheatley, P. S.; Zhao, X. B.; Fletcher, A. J.; Fox, S.; Rossi,
of crystalline porous materials. Micropor. Mesopor. Mater. 2012, A. G.; Megson, I. L.; Bordiga, S.; Regli, L.; Thomas, K. M. et al.
149, 134–141. High-capacity hydrogen and nitric oxide adsorption and storage in a
[55] Martin, R. L.; Smit, B.; Haranczyk, M. Addressing challenges of metal-organic framework. J. Am. Chem. Soc. 2007, 129, 1203–1209.

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research

You might also like