You are on page 1of 11

pubs.acs.

org/crystal Article

Mechanical Properties of a New Hybrid Inorganic−Organic


Framework: A Nanoindentation, High-Pressure X‑ray Diffraction,
and Computational Study
Li-Jun Ji, Zhi-Gang Li, Li-Yuan Dong, Ying Zhang, Tian-Meng Guo, Fei-Fei Gao, Guo-Qiang Feng,*
and Wei Li*
Cite This: Cryst. Growth Des. 2022, 22, 6984−6994 Read Online
Downloaded via NORTHWESTERN POLYTECHNICAL UNIV on March 2, 2023 at 16:54:35 (UTC).
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ACCESS Metrics & More Article Recommendations *


sı Supporting Information

ABSTRACT: The comprehensive mechanical properties of a new


synthesized 3D hybrid inorganic−organic framework (TPrA)-
Cu2(dca)5 (TPrA+ = N(C3H7)4+, tetrapropylammonium; dca− =
N(CN)2−, dicyanamide) have been studied via experimental
approaches and first-principles calculations. The nanoindentation
test demonstrates that Young’s moduli (E) and the hardness within
the 2D [Cu2(dca)4] layers are respectively 65 and 70% larger than
the direction normal to the 2D planes, and the framework of
(TPrA)Cu2(dca)5 is prone to cleavage along the 2D [Cu2(dca)4]
layers to lead to discrete displacement bursts in the initial loading
part of the load-indentation depth (P-h) curve of the <020>
direction. The bulk modulus of (TPrA)Cu2(dca)5 is 6.97 GPa within a pressure scope of 0 to 3.36 GPa, which is comparable to
those from porous MIL-47 and ZIF-8. The high anisotropy of Young’s moduli of 31.8 and the shear moduli of 45.0 provided via first-
principles calculations are an order of magnitude larger than those from many known porous and dense frameworks but close to
those of 2D hybrid systems. The broad range of Poisson’s ratio of (TPrA)Cu2(dca)5 indicates its very anisotropic response behavior
when under the uniaxial and shearing stress. The results of nanoindentation, synchrotron high-pressure X-ray diffraction, and first-
principles calculations synergistically indicate that the 3D architecture of (TPrA)Cu2(dca)5 has the potential to cleavage into 2D
nanosheets under the uniaxial or shearing stress. Further high-resolution microscopic characterization directly confirms the
successful exfoliation of the 3D framework of (TPrA)Cu2(dca)5 into 2D nanosheets via simple surfactant-free solvent-mediated
sonication and demonstrates that the (−102) plane is the cleavage plane.

■ INTRODUCTION
Hybrid inorganic−organic framework materials are a family of
multiferroic, optical, and magnetic properties in functional
devices which are traditionally dominated by inorganic or
single-phase compounds in the crystalline state, which are organic materials.5
formed by linking inorganic metal (M) clusters or ions with Although there have been a great large number of publications
multitopic organic bridges (L) to construct infinite arrange- aiming at designing new architectures and developing novel
ments in an M-L-M connected manner in at least one functions,2,5 only a small amount of studies have been aimed at
dimension.1 Over the past two decades, this family has drawn understanding the mechanical behaviors which deserve more
increasing attention of researchers by their enormous chemical attention. Because on the one hand, the framework structures
and structural diversity, and a large number of researchers have have to possess sufficient stiffness and resilience under the
focused on creating exotic properties or combining unique mechanical loading to refrain from distortion, collapse, and
properties that existed in purely inorganic or organic amorphization. On the other hand, functions such as
compounds by designing and synthesizing new structures.2 piezoluminescence and piezoelectricity are strongly dependent
Now to date, the researchers have usually sorted the hybrid on the elastic properties of the architectures.6 Motivated by the
frameworks into two main categories. One is nanoporous
coordination polymers known as metal−organic frameworks
(MOFs).3 Their novel structural natures such as ordered pore Received: June 13, 2022
structure, tailorable pore size, large specific surface area, and Revised: October 27, 2022
ultrahigh porosity endow them with significant potential in Published: November 7, 2022
many practical applications.4 The other is dense hybrid
frameworks, the structures of which are more similar to those
of classical inorganic materials, showing semiconductivity and

© 2022 American Chemical Society https://doi.org/10.1021/acs.cgd.2c00658


6984 Cryst. Growth Des. 2022, 22, 6984−6994
Crystal Growth & Design pubs.acs.org/crystal Article

ultimate goal of applying these promising frameworks into 3H2O (2 mmol, 0.484 g) was dissolved in 15 mL of water and combined
practical applications, the mechanical response behavior of with a mixture liquid of Na(dca) (dca− = N(CN)2−, dicyanamide) (6
many representative porous frameworks (i.e., ZIF-8 and MOF- mmol, 0.534 g) and (TPrA)Br (TPrA+ = N(C3H7)4+, tetrapropylam-
5) and layered molecular crystals, Omeprazole for instance, has monium) (2 mmol, 0.532 g) in 15 mL water. Dark green crystals
appeared by slow evaporation after 3 weeks. Then the crystals were
been extensively studied via nanoindentation and first-principles manually picked up from the liquid, and their purity was validated by
calculations, and the important information including the powder X-ray diffraction (Figure S1). We collected the infrared spectra
elasticity and plasticity, hardening and stiffness, and structural in the range of 4000−400 cm−1 via a Bruker Vertex 70 FTIR
evolution under stress has been reported.7−14 Furthermore, how spectrometer equipment. FT-IR (KBr): ṽ = 3754 (w), 3659 (w), 3435
weak interactions and microscopic structural topology work on (m), 3124 (w), 3087 (w), 2979 (w), 2942 (w), 2884 (w), 2781 (w),
the dynamic phenomena like the phase transition, phase 2398 (w), 2337 (m), 2311 (s), 2279 (s), 2195 (vs), 1633 (w), 1479
deformation, and elastic flexibility of 2D organic crystals have (w), 1458 (w), 1393 (m), 1352 (m), 1160 (w), 1101 (w), 1053 (w),
been comprehensively studied.15,16 For instance, the π-stacking 1031 (w), 1002 (w), 967 (w), 938 (w), 843 (w), 761 (w), 649 (w), 505
and C−H···O belonging to weak interaction play a main role in (m), 422 (w) cm−1. Elemental analysis was performed by using a Vario
Micro Cube elementar. The elemental analysis calcd (%) for
the loading-caused phase transition between the harder caffeine (TPrA)Cu2(dca)5 (Mr = 321.83): C 41.05, H 4.38, N 34.82; found:
and the softer 4-chloro-3-nitrobenzoic acid.15 Moreover, C 41.02, H 4.19, N 34.74.
significant mechanical responses like brittle fracture, plastic Single-Crystal X-ray Diffraction. We used a Rigaku XtalAB mini
bending, plastic shearing, elastic flexibility, and elastic bending diffractometer which was equipped with graphite monochromated Mo
under mechanical deformation tests of layered organic crystals Kα (λ = 0.71073 Å) to collect the single-crystal X-ray diffraction data. A
are investigated.17−20 However, studies concerning the stress crystal with a dimension of 0.78 × 0.42 × 0.21 mm3 was affixed on a
and strain related behavior of Cu-based frameworks are few, the glass fiber by glue and collected over the range of 2.606 to 29.579° (θ)
structure−mechanical property relationship is not comprehen- at 300 K. Data processing including reduction, face indexation, and
sive, and controlling and directing the mechanical behavior of absorption corrections was performed by the CrysAlisPro33 program.
The structure was solved by the direct method and refined based on F2
hybrid frameworks by using inorganic or organic building blocks by the full matrix least-squares method using a SHELX-9734 software
with different stiffness is still an issue.21,22 package. All nonhydrogen atoms were refined with anisotropic thermal
Hitherto, the two main categories of methods are widely parameters. The positions of hydrogen atoms attached to nitrogen
implied to prepare nanosheets for the MOF family. The first is a atoms and carbon atoms were generated geometrically using a riding
top-down route referring to peeling the layers from their mother model. The Olex235,36 software was used as an interface for the SHELX
bulks. Owing to the stiff covalent bonds within the layer and program. The crystallographic parameters are summarized in Table S2,
fragile interactions among layers, the layered bulk MOFs (e.g., the length of selected bonds and the selected angles are listed in Table
[Fe(bimX)2] (HBIMX = benzimidazole functionalized in the 5- S3, and the asymmetric unit is shown in Figure S2. Crystallographic
position, X = H, Cl, Br, CH3, or NH2)) can be cleaved into data for the compound have been deposited at the Cambridge
Crystallographic Data Centre (Deposit no. 2083249) and can be
nanosheets by approaches such as Scotchtape,23,24 sonica- obtained free of charge via www.ccdc.cam.ac.uk/data_request/cif.
tion,25,26 intercalation,27,28 milling,29,30 and freeze−thaw.31,32 Nanoindentation. The single crystals possessing a completeness
The second is a bottom-up route which means synthesizing the level of over 93% were selected via single-crystal X-ray diffraction by fast
MOF nanosheet by assembling the metal ions with organic collecting a sphere of data using an incremental scan method. Then
ligands. Both methods have long been considered and only can before doing the nanoindentation, the selected single crystal was firmly
be used to generate nanosheets from MOFs which originally affixed on iron stages. The experiments were operated on two natural
possess a layered structure. To further broaden the range of faces (010) and (−102) (Figure S3) by using a Berkovich diamond
available precursors and enrich their corresponding applications, nanoindenter with a three-sided pyramidal shape (Triboindenter of
trying 3D MOFs as exfoliation sources is also a feasible way. Hysitron, Minneapolis, USA). The load (P) and displacement (h) of
tested single crystals were continuously monitored and recorded by the
Herein, we concentrate on the mechanical properties of a new equipment with a force resolution of 5 mN and a displacement
synthesized 3D hybrid framework (TPrA)Cu2(dca)5. Specifi- resolution of 0.2 nm. The tip radius of the indenter applied to indent the
cally, we probed the elastic properties, plastic properties, and crystals was about 100 nm. To access regions which are flat enough for
high-pressure behaviors of this hybrid framework by nano- the experiment, we imaging the crystal surfaces by the same indenter tip
indentation, first-principles calculation, and synchrotron high- in AFM mode before doing the nanoindentation.37 The 10 × 10 μm2
pressure X-ray diffraction (HP-XRD) approaches. Our results regions with root mean square roughness (Rq, RMS roughness) below
give that the outstanding anisotropy of the mechanical 30 nm were selected before each loading. Rq and average roughness Ra
properties of this Cu-based 3D framework can be attributed to were 6.8 and 6.2 nm respectively for the selected region in the (020)
the synergy of the rigid 2D [Cu2(dca)4] connectivity along the plane (Figure 2b), and 15.8 and 8.9 nm respectively for the selected
region in the (−102) face (Figure 2c). The loading and unloading rates
<−102> direction and the sparse M-L-M bridges along <20−3> were 0.5 mN/s, and the hold time at peak load was 10 s. We performed
directions which connect adjacent [Cu2(dca)4] layers to form a seven indentations on each crystallographic face (Figure S4). To avoid
3D architecture. Furthermore, scanning electron microscopy the elastic recovery of the residual impressions after indentation, we
(SEM), transmission electron microscopy (TEM), and atomic captured the indentation impressions immediately after unloading by
force microscopy (AFM) characterization reveal that because AFM mode.
this 3D dense framework possesses prominent mechanical We used quasi-static mode to conduct the nanoindentation
anisotropy which is even much larger than some classic 2D experiments. The unloading stiffness (S) is established by the equation
materials (Table S1), the (TPrA)Cu2(dca)5 can be piled into S = dP/dh at the initial point of unloading. Therefore, we can find
nanosheets via simple surfactant-free solvent-mediated soni- Young’s modulus (E) and hardness (H) at the maximum depth of
penetration. As the fact that both the specimen and the indenter
cation. undergo elastic deformation, according to eq 1, the reduced modulus

■ EXPERIMENTAL SECTION
Synthesis. All chemicals and solvents were purchased with reagent
(Er) was used in analysis to explain it.

Er =
2
S
Ac
grade and used as received. In the synthesis procedure, Cu(NO3)2· (1)

6985 https://doi.org/10.1021/acs.cgd.2c00658
Cryst. Growth Des. 2022, 22, 6984−6994
Crystal Growth & Design pubs.acs.org/crystal Article

Figure 1. Crystal structure of (TPrA)Cu2(dca)5 viewed along the b-axis (a) and c-axis (b), where TPrA+ is tetrapropylammonium and dca− is
dicyanamide. Color scheme: Cu2+, olive; C, light gray in dca− ligand and dark gray in TPrA+ cation; N, blue. The TPrA+ cations in (a) and hydrogen
atoms in (b) are omitted for clarity.

In eq 1, the Ac is the projected contact area under load, and β is a integrated by the FIT2D software package.42 The lattice parameters of
constant which depends on the geometry of the indenter (β = 1.034 for the unit cell obtained from Le Bail fits are listed in Table S4.
the Berkovich tip). Based on the Oliver−Pharr method,38,39 for a given First-Principles Calculation. Cambridge Sequential Total Energy
sample, its Young’s modulus can be extracted from the reduced Package (CASTEP)43,44 was used to perform the first-principles
modulus according to eq 2: calculations. This energy package is based on plane-wave pseudo-
potential density functional theory (DFT).44 We developed the
1 1 vi2 1 vs2 function by using Perdew, Burke, and Ernzerhof in the generalized
= +
Er Ei Es (2) gradient approximation (PBE-GGA)45 form to represent the exchange-
correlation energy. We employed the finite method to determine the
In eq 2, E and v are Young’s modulus and Poisson’s ratio, respectively. elastic constants of (TPrA)Cu2(dca)5.46 The energy cut-off was fixed at
The subscripts i and s refer to the indenter and test material, 400 eV, and the k-point mesh sampling scheme of Monkhorst-Pack was
respectively. The indenter properties used in this study are Ei = 1141 set to 2 × 2 × 1 for geometry optimization, density of states, and elastic
GPa and vi = 0.07. The different faces of the single crystals were constants.47 The crystal structure was optimized at 0 GPa through the
assumed to possess isotropic elastic properties with vs = 0.3. The relaxation of both the lattice parameters and atomic positions. In
hardness of a material is defined as its resistance to local plastic geometry optimization, the convergence criteria for energy, maximum
deformation, and indentation hardness (H) thus can be determined stress, maximum displacement, and maximum force were set as 10−5 eV
from the maximum indentation load (Pmax) and can be divided by the per atom, 0.05 GPa, 10−4 Å, and 0.005 eV Å−1, respectively. After
contact area (Ac) according to eq 3: obtaining the elastic constants, the 3D distribution of Young’s modulus,
Poisson’s ratio, and shear modulus were calculated by the ELATE
Pmax
H= program.48
Ac (3) Scanning Electron Microscopy. The nanosheets were exfoliated
by dispersing 1 mg powder of (TPrA)Cu2(dca)5 in 1 mL of isopropanol
In eq 3, the contact area (Ac) is a function of the contact depth (hc), and then sonicated for about 90 min (120 W, 40 kHz). The milky
Ac = f(hc). The contact depth can be determined from the P−h data by suspension was centrifuged with a rate of 6000 r/min. The bottle
using eq 4: sediment including exfoliated nanosheets and incomplete exfoliated
Pmax bulks was dipped on a preheated clean silicon substrate. A layer of
hc = hmax conductive gold particles was coated on the surface of the sample after
S (4)
the solvent completely evaporated. The sample then was imaged by
where hmax is the maximum indentation depth and ε is a constant which field-emission scanning electron microscopy (FESEM, Sigma 500,
depends on the geometry of the indenter. Germany).
High-Pressure Synchrotron Powder X-ray Diffraction. We Atomic Force Microscopy. We used the Digital Instrument
performed the synchrotron HP-XRD characterizations by using the 4 Multimode SPM equipped with a BRUKER RTESPA-300 tip in the
W2 beamline in Beijing Synchrotron Radiation Facility (BSRF). The tapping mode to image the surface topographies of samples and offer
wavelength of the X-ray beam is 0.61990 Å, and the X-ray beam was the height profiles of nanosheets. The nanosheets were fabricated by
focused into a spot with a size of 36 × 12 μm2 by Kirkpatrick−Baez dispersing 0.1 mg of the (TPrA)Cu2(dca)5 powder in 5 mL of
mirrors. We used the diamond anvil cells (DACs) to generate a isopropanol (20 mL cylindrical vial) and sonicated for about 90 min in a
hydrostatic pressure, and their culet diameter is 400 μm. With the the KQ3200DE ultrasonic cleaner. The power and frequency were fixed
increase of pressure from atmospheric pressure, all the diffraction peaks at 120 W and 40 kHz, respectively. After the obtained milky suspension
move to high angles gradually, indicating the shrinkage of crystal was centrifuged with a rate of 4500 r/min, the vial was then kept still for
lattices. Nevertheless, when the pressure continually increased over the 5 min, dipping the top suspension on a clean preheated silicon substrate
threshold of 3.358 GPa, some weak diffraction peaks began to (1 × 1 cm2). The as-prepared nanosheets were scanned with a rate of
disappear, which manifested the partial amorphous process of the 0.80 Hz when the residual solvent on the substrate was evaporated. The
(TPrA)Cu2(dca)5 structure. To maintain the structural crystallinity, we resolution of the images was 512 × 512 pixel2. The operation point, P
analyzed the high-pressure behaviors of (TPrA)Cu2(dca)5 in the range gain, and integral gain were set as 0.153 V, 0.001, and 1000, respectively.
of 0∼3.358 GPa. Crystals of (TPrA)Cu2(dca)5 were sufficiently ground Transmission Electron Microscopy. The nanosheets used in this
into powder and subsequently loaded with a few ruby balls in a cave of experiment were prepared by dispersing 0.1 mg powder of (TPrA)-
about 200 μm diameter in a preindented stainless steel gasket with a Cu2(dca)5 in 1 mL of isopropanol. The above powder was sonicated for
thickness of about 40 μm.40 Both the sample and ruby balls were sealed about 90 min (120 W, 40 kHz). Droplets of the suspension were dipped
in DACs with the silicone oil as the pressure-transmitting media at onto the holey carbon-coated Cu grids and evaporated in the air. We
room temperature. We calibrated the pressure in DACs by measuring operated the Titan 60−300 Cs Corrected transmission electron
the fluorescence of the ruby balls.41 The Pilatus 2 M detector was used microscope at 300 kV to capture the bright-field and high-resolution
to collect the diffraction patterns. The diffraction patterns then were TEM (HRTEM) images of nanosheets and used the JEM-2100

6986 https://doi.org/10.1021/acs.cgd.2c00658
Cryst. Growth Des. 2022, 22, 6984−6994
Crystal Growth & Design pubs.acs.org/crystal Article

Figure 2. (a) P−h curves normal to the (020) and (−102) planes of single crystals of (TPrA)Cu2(dca)5 measured with a Berkovich tip. (b, c) Typical
AFM height topology obtained from the residual indents corresponding to the P−h curves depicted in (a). Rq and Ra were 6.8 and 6.2 nm for the region
in (b) and 15.8 and 8.9 nm for the region in (c). Their cross-sectional profiles along the three designated directions are plotted in (d, e).

electron microscope at 200 kV to obtain the selected area electron N7−C16 being about 119.5° and the distance of Cu-dca−-Cu
diffraction (SAED) pattern, because the nanosheets of (TPrA)- connectivity is about 0.71 nm, while the dca− linkers between
Cu2(dca)5 were frail and easy to lose crystallinity within 15 s when adjacent 2D [Cu2(dca)4] layers are stretched with an angle of
under the electron beam. C20−N16−C22 of about 124.2° and the distance of Cu-dca−-

■ RESULTS AND DISCUSSION


Crystal Structure. (TPrA)Cu2(dca)5 (TPrA+ = N(C3H7)4+,
Cu connectivity is about 0.83 nm. The inspection above
indicates that the bonding strength within the 2D [Cu2(dca)4]
plane is much stronger than it is between [Cu2(dca)4] planes.
tetrapropylammonium; dca− = N(CN)2−, dicyanamide) crys- Moreover, the (−102) plane which is coincident with the 2D
tallizes in the orthorhombic space group Pbca with cell [Cu2(dca)4] layer is the morphologically major face with the
parameters a = 13.9367(8), b = 16.4222(9), and c = largest interplanar spacing. Therefore, the (−102) has the lowest
25.878(2) Å. The asymmetric unit of the structure is composed attachment energy which is calculated via Bravais−Friedel−
of two Cu2+, five dca− ligands, and a TPrA+ guest. The Cu2+ ion Donnay−Harker (BFDH) theory.52 Hence, (−102) can be
is five-coordinated in a pyramidal geometry by five dca− ligands. considered as a facile slip plane for allowing intermittent flake
The Cu−N distances are in the range of 1.970(3)−2.258(3) Å, movement along the b-axis during mechanical loading. For most
and the N−Cu−N angles lie in the range of 88.5(7)−174.7(3)° 2D structures, layer stacking via weak interactions is known to
(Figure S2, Tables S2−3). Each Cu1N5 polyhedron links two allow sudden and collective slip to relax and accommodate the
Cu1N 5 polyhedrons along the b-axis and one Cu2N 5 loading stress imposed on the mechanically deformed
polyhedron along <10−2> direction via three single dca− crystals.15,16,20
linkers but links one Cu1N5 polyhedron along the a-axis via Nanoindentation Study. To elucidate the structural
double dca− ligands, while each Cu2N5 polyhedron bridges two stiffness of the 3D hybrid framework (TPrA)Cu2(dca)5, we
Cu2N5 polyhedrons along the b-axis by linking two single dca− carried out nanoindentation measurements on the naturally
linkers and one Cu2N5 polyhedron along <10−2> direction by grown (020) and (−102) faces of the single crystals using a
one single dca− ligand. The inorganic CuN5 polyhedrons and sharp Berkovich tip in the quasi-static mode, and the maximum
organic dca− linkers corporately form the infinite [Cu2(dca)4] load was set as 5 mN.49,50 The naturally grown two facets agree
connectivity, and this connectivity extends approximately within well with the BFDH morphology calculated from Mercury
the (−102) plane to form a corrugated 2D layer. The adjacent software. The representative load-indentation depth (P−h)
2D [Cu2(dca)4] layers are about 0.6 nm apart and bridged by the curves obtained on the two different faces are shown in Figures
tilt Cu1-dca−-Cu2 connectivity with a length of about 0.82 nm 2a and S4.
along the <20−3> direction to produce a 3D [Cu2(dca)5]− The P−h curves accomplished on the two natural facets are
framework (Figure 1a). This anion framework is charge different and exhibit significant residual depths upon complete
balanced by the TPrA+ guest cations which are located in the unloading, indicating that the two planes undergo significant
framework cavities, and the projections of the TPrA+ cations plastic deformation during the indentation. The (020) face
along the b-axis are largely overlapped (Figure 1b). Further shows a smaller penetration depth (hmax) of about 478 nm as
examination suggests that the dca− ligands within 2D compared to the (−102) face with a hmax of about 637 nm. The
[Cu2(dca)4] layers are highly flexed with the angle of C13− high value of hmax for (−102) correlates well with its observed
6987 https://doi.org/10.1021/acs.cgd.2c00658
Cryst. Growth Des. 2022, 22, 6984−6994
Crystal Growth & Design pubs.acs.org/crystal Article

high plastic behavior and demonstrates its smaller hardness up merely on one side for the (020) face.55 It is noted that the
(Figure 2c). extensive pile-ups around the edge of the indentation can result
The loading segment of the P−h curve obtained on (−102) is in the underestimation of the contact area (Ac) for isotropic
smooth, presenting a homogeneous plastic deformation. In materials; therefore, Young’s modulus and hardness were
contrast, the loading part of the P−h curve on (020) presents overestimated. For anisotropic crystals, researchers try to read
discrete displacement bursts (“pop-ins”). Such discrete events Ac directly from the residual indents at a fully unloaded state to
signify intermittent plasticity instead of a continuous change. measure the degree of anisotropy more precisely. Then they
The magnitudes of the displacement burst (hpop‑in) were found found that the recalculated E and H are still relatively lower
to be 5.7, 8.1, 6.3, and 7.9 nm, respectively, which are 7 to 10 because of pile-ups, but the anisotropic ratios of E and H are
times interplanar d-spacing of the (020) facet (∼0.8 nm).51 The comparable with those obtained using the Oliver−Pharr
“pop-ins” originate from the collective and sudden sliding of method.21,50
multiple (−102) planes during indentation, because the 2D The average values of Young’s modulus (E) and hardness (H)
[Cu2(dca)4] layers are oriented approximately parallel to the extracted from the P−h curves (Figure 2a) are 15.18(29) and
indenter axis (b-axis) when indenting the (020) face, and the 0.46(03) GPa for the (002) face, and 9.21(18) and 0.27(28)
Cu2−N15−C22−N16−C20−N12−Cu2 linkages which work GPa for the (−102) face, respectively, by using the Olive−Pharr
for binding the 2D layers together along the <20−3> direction method in which the pile-up effects are not accounted for. The
are broken although the strength of most coordination bonding ratio53,54 between the two planes in E and H is E(002):E(−102) =
typically is of about 50 to 200 kJ mol−1, the shearing of the layers 1.65:1 and H(002):H(−102) = 1.7:1, respectively, giving a change in
consequently occurs to produce the bursts of displacements the modulus of about 65%. The (020) face was found to have a
(Figure 1a). In contrast, no “pop-ins” were observed on the higher modulus, while the (−102)-oriented face was found to be
(−102) face when loading along the normal orientation of 2D of lower modulus, indicating that the 2D [Cu2(dca)4] layers
layers, such a phenomenon can be construed as the shrinkage of which are bonded via covalent Cu2−N15−C22−N16−C20−
soft organic (TPrA)+ cations which act as spacers to shorten the N12−Cu2 chains are relatively compliant when loaded along the
interlayer distance persistently. normal orientation ((−102) direction). According to the crystal
It is worth noting that the occurrence of “pop-ins” for 2D structure of the (TPrA)Cu 2 (dca)5 framework, the 2D
hybrid frameworks is quite different from that in layered organic [Cu2(dca)4] layers are connected via a single type of chain
crystals. For layered organic crystals, the “pop-ins” generally (Cu2−N15−C22−N16−C20−N12−Cu2) along the <20−3>
appear and are more dominant in the direction normal to the direction, and the length and the bending angle of this type of
layers because of the microscopic elastic compression of the chain are 0.83 nm and 124.2°, respectively. In contrast, the Cu2+
layers followed by a sudden slippage upon reaching a critical load cations within layers are connected via three types of chains
applied by the indenter tip.15,16,20 In contrast, for hybrid and (double Cu1-N6-C16-N7-C13-C2-Cu1 chains along the a-axis,
inorganic structures, the indenter axis that is approximately two Cu1-N3-C14-N5-C15-N4-Cu1 chains along the b-axis, and
parallel to the layers often leads to obvious or more significant single Cu1-N8-C17-N9-C18-N10-Cu2 along the <−102>
displacement bursts than that normal to the layers. It arises from direction), and the lengths and bending angles of these three
the breakage of weak bonds between neighboring layers such as types of the chain are about 7.1 nm and 119.5° for Cu1−N6−
van der Waals forces, hydrogen bonds, and covalent bonds C16−N7−C13−C2−Cu1, 0.81 nm and 122.1° for Cu1−N3−
inducing shearing of layers.21 Such a correspondence was also C14−N5−C15−N4−Cu1, and 0.82 nm and 123.1° for Cu1−
witnessed in 2D hybrid crystals such as Mn(C6H8O4)(H2O), N8−C17−N9−C18−N10−Cu2. It is obvious that the lengths
Cu3(H2O)2[O3PCH2CO2]2, and (TPrA)Cu(NO3)(dca)2.21,55 and bending angles of Cu-dca−-Cu chains within [Cu2(dca)4]
The AFM topography and cross-sectional profiles of indents layers are smaller than that between layers. The chain-like ligand
obtained from the (020) and (−102) faces are depicted in with a shorter length and a smaller bending angle generally
Figure 2. Only one minor (about 165 nm above the original possess stronger bonding strength and vice versa. Therefore, the
surface) pile-up appears on one side of the residue impression interlayer Cu-dca−-Cu chains are relatively weak and present
for the (020) plane (Figure 2b), while the more distinct “pile- lower resistance against external compressive and shear stresses.
ups” (about 200, 286, and 495 nm above the original surface) In contrast, the (020) face shows a higher modulus with a
can be seen on all three edges of the indent for (−102) face by magnitude about twice than that of the (−102) face. This is
applying the same load (Figure 2c). It can be seen that the pile- because the indenter orientation for the (020) experiments is
up behavior is highly anisotropic, indicating the underlying 2D approximately parallel to the 2D [Cu2(dca)4] layers, demon-
layered structure and the quite distinct interactions between strating that the Cu-dca−-Cu chains within 2D planes are stiffer
intra- and interlayer. According to the microscopic crystal and able to withstand a certain extent of elastic strains such as
structure nature of the (TPrA)Cu2(dca)5 framework, the translational and torsional motions of polyhedra, bond
[Cu2(dca)4] layers are broken plastically under the sustain stretching and bending, or combination of the above.50
increased loading stress normal to the (−102) plane, but the High-Pressure Properties. The nanoindentation gives that
broken fragments are hard to slip apart because of the for the 3D (TPrA)Cu2(dca)5 framework, the bonds within 2D
comprehensive obstruction of the Cu2−N15−C22−N16− [Cu2(dca)4] layers are significantly stiffer than that between
C20−N12−Cu2 covalent connectivity between layers, the layers when they are under the external elastic strains. Thus, the
high overlap of TPrA+ cations within the (−102) plane, and intralayer should be more compressible than the interlayer. To
the electrostatic force between TPrA+ cations and the quantitatively check this orientation-dependent compressibility
[Cu2(dca)5]− framework. The significant amount of broken of (TPrA)Cu2(dca)5, the high-pressure synchrotron X-ray
cleavage plane ejects out of the edges around the indenter, diffraction measurements were performed under the hydrostatic
therefore, leading to the dramatic pile-ups of the (−102) plane. environment. The representative diffraction patterns collected
In contrast, the oblique angle of the indentation direction from at 0, 1.82, and 3.36 GPa are shown in Figure S5a. It is clear that
the 2D [Cu2(dca)4] plane is responsible for the existence of pile- with the increase of pressure, no new diffraction peaks appeared
6988 https://doi.org/10.1021/acs.cgd.2c00658
Cryst. Growth Des. 2022, 22, 6984−6994
Crystal Growth & Design pubs.acs.org/crystal Article

during the compression process, and all the peaks move to high cavities of the anion [Cu2(dca)5]− framework. This value is
angles gradually, demonstrating that there was no phase significantly smaller than that of the dense stacked (Gua)[Zn-
transition, and the shrinkage of crystal lattices occurred due to (HCOO)3] (calc. 38.41 GPa)58 and metal-free (C4N2H12)-
the shortened interatomic distances. Although some weak [NH4I3·H2O]59 but comparable to that of porous MIL-47 (calc.
diffraction peaks disappear, the crystallinity of the sample is 9.74 GPa),60 ZIF-8 (calc. 9.21 GPa),9 and MOF-5 (calc. 17.61
retained in the whole pressure range from ambient to 3.36 GPa),7 certifying its much softer nature in resistance to
GPa,40,56 which is in significant contrast with the porous MOFs hydrostatic compression.
that have a relatively low amorphization pressure threshold.10 First-Principles Calculations. To further understand the
Based on the cell parameters extracted from Le Bail fits at each comprehensive elastic properties of the framework (TPrA)-
pressure (Table S3, Figure S5b), the compressibility of Cu2(dca)5, its full elastic tensors were calculated based on the
interplanar spacings of (−102), (100), and (020) facets is optimized crystal structure at 0 GPa by using the DFT-GGA +
calculated via the Bragg diffraction formula with a magnitude of PBE method.61,62 For the orthorhombic system, the obtained
about 9.7, 7.1, and 9.1%, respectively, within a pressure scope of nine independent elastic constants and derived Young’s moduli
3.36 GPa. The maximum change in compressibility is about (E), shear moduli (G), and Poisson’s ratios (v) are listed in
37%, and it is clear that the (100) plane is the toughest. As Table 1. The theoretical values of E(002) and E(−102) extracted
presented in Figure 1a, the Cu1N5 polyhedron links its nearest from the above elastic
neighbor Cu1N5 via double Cu1−N6−C16−N7−C13−C2− constants are 12.9 and 10.8 GPa, respectively, which is in good
Cu1 chains with the shortest chain length (7.1 nm) and smallest agreement with the aforementioned nanoindentation results,
chain bending angle (119.5°) to form unbroken loop-like testifying to the rationality of the DFT elastic tensors. The bulk
linkages oriented along the <100> direction. Such structural modulus extracted from the matrix is 12.71 GPa, which is 5.73
nature not only adds extra stiffness to the 2D [Cu2(dca)4] layers, GPa larger than the experimental synchrotron powder
but also leads to further anisotropy. As external hydrostatic diffraction value of 6.97 GPa. The two following factors are
pressure was applied, the (100) facet was compressed along the response for the difference: (i) The DFT calculations were
<100> direction, where the smallest compressibility was performed at 0 K, while the high-pressure synchrotron powder
obtained. By contrast, the (−102) and (020) faces have diffraction experiments were performed under ambient
relatively low compressibility. In contrast, the 2D [Cu2(dca)4] conditions.63 (ii) The silicone oil solidifies at 2.0 GPa but
layers lie in the (−102) plane which is bounded via the single remains nearly quasi-hydrostatic in the pressure range which
would affect the experimental results.64,65 Accordingly, the
Cu2−N15−C22−N16−C20−N12−Cu2 chain with the longest
larger calculated values are expected, and the 1−6 GPa
chain length (0.83 nm) and largest chain bending angle
difference is reasonable and common.40,65−67
(124.2°), and the (−102) face thus is more compliant and
In terms of Young’s modulus reflecting the ability of a given
presents largest compressibility. As depicted in Figure 1b, the
material to resist uniaxial deformation, the highly distorted 3D
Cu1N5 polyhedron within 2D [Cu2(dca)4] layers extends the
surface reflecting significant anisotropy is shown in Figure 4a,d.
two Cu1−N3−C14−N5−C15−N4−Cu1 chains along <020> E reaches the maxima along the <805> direction with a
and <0−20> directions, respectively, to bridge the Cu1N5 with a magnitude of 14.6 GPa, and this direction approximately lies in
middle chain length (0.81 nm) and chain bending angle the (−102) plane which is along the rigid 2D [Cu2(dca)4]
(122.1°), explaining why the compressibility of (020) was found connectivity plane. Thus this orientation exhibits the largest
to lie between those of the (−102) and (100). Moreover, we also stiffness. In addition, E reduces to the minima along the <011>
discuss the bulk modulus of the (TPrA)Cu2(dca)5, and the unit direction with the value of 0.5 GPa, which is attributed to the
cell volume decreases by 24.9% at 3.36 GPa (Figure 3). The collective influences including the staking manner of the 2D
pressure−cell volume data were fitted by using the PASCal [Cu2(dca)4] layer along the <020> direction via sparse and soft
software with the second-order Birch−Murnaghan equation.57 dca− ligands and the staggered arrangement of soft TPrA+
The fitted bulk modulus is 6.97 GPa lying in the range of values cations along the c-axis. Notably, the anisotropy of E of
for many known porous MOFs (∼5−30 GPa)53 but adopted a (TPrA)Cu2(dca)5 defined as the ratio AE = Emax/Emin is about
nonporous structure with the large TPrA+ cation located in the 31.8, which is an order of magnitude larger than that of both
common porous frameworks like ZIF-8 (1.4)9 and MOF-5
(4.3)7 and dense framework like (C4N2H12)[NH4I3·H2O]
(2.2)59 and (Gua)[Zn(HCOO)3] (2.7),58 even significantly
larger than those of typical 2D materials, such as graphene
(28.4),68 black phosphorus (6.5),69 MoS2 (4.6),70 and
(Benzylammonium)2PbBr4 (4.9)71 (Table S1). As the out-
standing anisotropy of Young’s modulus always indicates an
extreme anisotropic bonding manner which is often seen in the
2D materials’ family, this feature of (TPrA)Cu 2 (dca) 5
accordingly indicates that this 3D hybrid framework could be
exfoliated into nanosheets like 2D structures via applying
uniaxial strain.
For a framework, the shear modulus reflects the resistant
ability to change in shape or angle. The 3D and 2D distribution
Figure 3. Evolution of unit cell volume as a function of pressure. The of shear modulus are presented in Figure 4b,e, respectively. The
experimental data with error bars are presented as dark gray dots, and G reaches the maximum value of 5.3 along the <−47−6>
the second-order Birch−Mumaghan fitting curve obtained by using direction when shearing the (−476) plane, while reaches the
PASCal is shown as a green line. minimum value of 0.1 along the <001> when shearing the (010)
6989 https://doi.org/10.1021/acs.cgd.2c00658
Cryst. Growth Des. 2022, 22, 6984−6994
Crystal Growth & Design pubs.acs.org/crystal Article

Table 1. Elastic Constants and the Maximal and Minimal Values of Elastic Properties of (TPrA)Cu2(dca)5 and Other Typical 3D
Crystalsa
Cij (GPa)
elastic properties C11 C22 C33 C44 C55 C66 C12 C13 C23
(TPrA)Cu2(dca)5 32.74 20.96 16.39 0.12 4.71 2.03 16.20 15.76 8.29
E (GPa) G (GPa) v
Emax Emin AE Gmax Gmin AG vmax vmin Av K (GPa)
14.6 0.5 31.8 5.3 0.1 45.0 2.7 −1.7 6.97b
ZIF-89 3.9 2.7 1.4 1.4 0.9 1.5 0.6 0.3 2.0 9.21
MOF-57 21.3 4.9 4.3 8.2 1.7 4.8 0.8 0.09 8.9 17.6
(C4N2H12)[NH4I3·H2O]59 53.3 24.0 2.2 21.4 7.5 2.8 0.6 0.1 6.9 29.47
(Gua)[Zn(HCOO)3]58 70.9 26.6 2.7 24.8 11.0 2.3 0.8 0.1 6.0 31.82
a
All extremum values were derived by using the ELATE48 software. Anisotropy of X is defined as Ax = Xmax/Xmin. bExtracted from the high-pressure
synchrotron diffraction experiment.

Figure 4. 3D and 2D plots of Young’s modulus E (a, d), shear modulus G (b, e), and Poisson’s ratio v (c, f) of (TPrA)Cu2(dca)5. In (b, c), the
transparent outer layer and nontransparent inner layer represent the maximum values and minimum values, respectively. In (e, f), the blue line and red
line correspond to the maximum and minimum values, respectively.

plane. These values give rise to a remarkable anisotropic index value of 1.75, demonstrating that its boundary nature is between
AG of 45, which is characterized as the ratio of Gmax/Gmin. This ductile and brittle.
anisotropic index is more than an order of magnitude higher Liquid Exfoliation. From the above analysis of the crystal
than it of both porous and dense frameworks listed in Table 1, structure−mechanical property relationship, we can know that
also significantly larger than it of prototypical 2D materials listed the 3D column-laminar framework (TPrA)Cu2(dca)5 is formed
in Table S1. This high shear anisotropy of (TPrA)Cu2(dca)5 via 2D [Cu2(dca)4] layers extending in the (−102) plane and 1D
indicates its structural instability under the shearing stress chain-like Cu2−N15−C22−N16−C20−N12−Cu2 linkages
although it is formed via strong coordinative and covalent bonds. used to bound the adjacent layers. As the Cu-dca−-Cu linkages
Turning attention to Poisson’s ratio (v), which is charac- of the intralayer show a shorter bonding distance and a smaller
terized as the ratio of the transverse strain (εj) to the axial strain bending angle than those of interlayer, the bonding strength of
2D [Cu2(dca)4] planes is significantly stronger than that
(εi), v = −εj/εi. As shown in Figure 4c,f, the Poisson’s ratio is
between layers. In this regard, it can be seen that the
maximized to 2.7 when stretched along the <924> direction,
nanoindentation and high-pressure behaviors are highly
giving rise to a lateral contraction along the <177> direction. anisotropic, demonstrating the underlying 2D layered structural
The Poisson’s ratio is minimized to −1.7, when loaded along the characteristics and the distinct interactions between the
<924> direction with a lateral expansion along <−233> intralayer and interlayer. For instance, the occurrence of “pop-
direction. Notably, the Poisson’s ratio ranges from −1.7 to ins” on the (020) facet and the orientation-dependent pile-up
2.7, which is much wider than that of many well-known 3D features in nanoindentation measurements collaboratively
porous and dense materials, lying on either side of the value of indicate that the (−102) plane is prone to generate slippage
0.26, implying its significant anisotropy and between ductile and when the applied uniaxial stress is oriented parallel to the 2D
brittle nature. This boundary feature is also verified by Pugh’s [Cu2(dca)4] planes. Further DFT calculations also indicate that
criterion,72 according to which the K/G ratio of (TPrA)- the anisotropy indexes are comparable to even larger than the
Cu2(dca)5 is in the range of 1.3−69.7, including the threshold 2D family, which manifests the structural instability of the 3D
6990 https://doi.org/10.1021/acs.cgd.2c00658
Cryst. Growth Des. 2022, 22, 6984−6994
Crystal Growth & Design pubs.acs.org/crystal Article

Figure 5. AFM surface topographies (a) and the corresponding height profiles (b) of multilayered nanosheets. TEM bright-field image (c) of ethanol
exfoliated nanosheets supported on a holey copper mesh. HRTEM image (d) of the 2D nanosheets showing the crystal lattice, the inset is the SAED
pattern.

(TPrA)Cu2(dca)5 framework. Taking those into account, we morphology of these nanosheets indicates incomplete peeling,
speculate that the (TPrA)Cu2(dca)5 could be exfoliated into 2D which could arise from the restacking of the exfoliated
nanosheets by simply applying uniaxial and shearing stress. As nanoflakes possibly through the strong electrostatic attractions.
the liquid phase exfoliation process contains both shearing and The HRTEM image was collected at the edge of the exfoliated
uniaxial stresses along any orientation, we exfoliated this 3D nanosheets and is displayed in Figure 5d. The lattice fringe
(TPrA)Cu2(dca)5 by simple surfactant-free solvent-mediated shows an interplanar distance of about 0.52 nm, which can be
sonication to confirm our conjecture. One milligram of well- associated with the (221) plane as it agrees well with its d-
ground powder of the bulk crystals was dispersed in 1 mL of spacing of 0.52 nm. More importantly, the SAED patterns were
isopropanol, then sonicating this mixture for a period of 90 min. collected normally to the nanosheets to verify their atomic-level
We performed SEM to examine the morphologies of these ordered structure. The diffracted spots in Figure 5d inset
products and get direct evidence of successful cleavage. As seen correspond to the (221) and (2−21) planes of the (TPrA)-
in Figure S6, all incompletely exfoliated bulks show stepwise Cu2(dca)5 nanosheets. This result associated with the HRTEM
edges with peeled nanosheets within 500 nm adhered on their image indicates that the (TPrA)Cu2(dca)5 was exfoliated
surfaces, confirming the successful exfoliation of (TPrA)- parallel to the (−102) plane and the (−102) is the slippery
Cu2(dca)5. As SEM only can offer information about the lateral facet, which verified our aforementioned conjecture and agrees
lengths of flakes, we further perform AFM to check the thickness well with our nanoindentation and high-pressure synchrotron X-
of the terraces. As seen in Figure S7, the incompletely exfoliated ray diffraction results.
bulks with different color contracts show terraces at their edges.
The height profile displays several steps which are approximately
5−10 nm in thickness (Figure S7). To observe the exfoliated
nanosheets more directly, we dilute the aforementioned
■ CONCLUSIONS
In summary, the mechanical properties of a new hybrid
suspension to 0.02 mg/mL to minimize agglomeration, and framework (TPrA)Cu2(dca)5 have been systematically inves-
then the suspensions were dipped on preheated Si wafers which tigated by experimental approaches and first-principles calcu-
were left to stand until the complete evaporation of the solvent. lations. Nanoindentation tests show that Young’s modulus of
As shown in Figure 5a,b, the exfoliated nanosheets with a lateral (−102) is 65% larger than that of the (002) plane, indicating a
size of about 341 nm agree with the SEM result, and the heights 1.65 bonding strength anisotropy between the two faces.
of steps measured in the height profile are 8.9, 13.6, 4.1, 13, 31.8, Furthermore, the discrete plastic deformations on the loading
59.3 and 12.3 nm, respectively. part of the (020) plane demonstrate that the 2D [Cu2(dca)4]
In addition, we performed TEM to investigate the crystalline planes have the least attachment energy and the 3D framework is
and check the cleavage direction of the sonication exfoliated prone to cleavage along the <−102> direction, namely, piled
nanosheets. As illustrated in Figure 5c, the bright-field image parallel to (−102) plane, to make displacement bursts in the P−
shows the flat (TPrA)Cu2(dca)5 nanosheets with various h curves. Inhomogeneous pile-ups around the indenter
dimensions stacked like overlapped leaves. The lateral distance impressions also verified this result. We have also probed the
of the nanosheets ranges from 200 to 500 nm. The stepwise hydrostatic properties of the (TPrA)Cu2(dca)5 via high-
6991 https://doi.org/10.1021/acs.cgd.2c00658
Cryst. Growth Des. 2022, 22, 6984−6994
Crystal Growth & Design pubs.acs.org/crystal Article

pressure synchrotron X-ray powder diffraction experiments, Authors


which were performed from ambient to 3.36 GPa. This study has Li-Jun Ji − Department of Physics and Mechanical & Electrical
demonstrated that the d-spacing of (−102), (100), and (020) Engineering, Hubei University of Education, Wuhan 430074,
shrank 9.7, 7.1, and 9.1% during the compression process, China; School of Materials Science and Engineering & Tianjin
showing the more rigid nature of the 2D [Cu2(dca)4] layer. Key Laboratory of Metal and Molecule-Based Material
Although (TPrA)Cu2(dca)5 adopts a nonporous structure, its Chemistry, Nankai University, Tianjin 300350, China;
bulk modulus calculated via the second-order Birch−Murna- orcid.org/0000-0001-9006-788X
ghan equation is 6.97 GPa lying in the range of values for porous Zhi-Gang Li − School of Materials Science and Engineering &
MOFs, indicating its much softer nature under hydrostatic state. Tianjin Key Laboratory of Metal and Molecule-Based Material
Furthermore, the full elastic constants of the framework Chemistry, Nankai University, Tianjin 300350, China
(TPrA)Cu2(dca)5 were calculated by first-principles calcula- Li-Yuan Dong − School of Science, Jiangsu University of Science
tions, and the result indicated that the anisotropy of Young’s and Technology, Zhenjiang 212100, China
moduli and shear moduli are an order of magnitude larger than Ying Zhang − School of Materials Science and Engineering &
both common porous and dense frameworks, even higher than Tianjin Key Laboratory of Metal and Molecule-Based Material
many typical 2D structures, indicating that this material is more Chemistry, Nankai University, Tianjin 300350, China
similar to 2D materials compared with 3D materials in response Tian-Meng Guo − School of Materials Science and Engineering
to uniaxial and shearing stresses. This feature demonstrates that & Tianjin Key Laboratory of Metal and Molecule-Based
the 3D framework (TPrA)Cu2(dca)5 could produce structure Material Chemistry, Nankai University, Tianjin 300350,
slippery and consequently be exfoliated into nanosheets like 2D China
structures under compression or stretching. The values of Fei-Fei Gao − School of Materials Science and Engineering &
Poisson’s ratio reflect that the boundary nature of (TPrA)- Tianjin Key Laboratory of Metal and Molecule-Based Material
Cu2(dca)5 is between brittle and ductile. Furthermore, SEM, Chemistry, Nankai University, Tianjin 300350, China
AFM, and TEM characterization gave direct evidence of the Complete contact information is available at:
successful exfoliation of (TPrA)Cu2(dca)5, and the HR-TEM https://pubs.acs.org/10.1021/acs.cgd.2c00658
image and SAED manifest that this 3D structure can be peeled
along to (−102) plane, which agree well with the nano- Funding
indentation results. Considering the soaring demand for
We acknowledge the financial support from the Guiding Project
searching for new 2D materials, this study could be instructive
of Education Scientific Research Plane in the Hubei Provincial
for selecting 3D structures which have extreme mechanical
Department (No. 82021257) and the National Natural Science
anisotropy and can be exfoliated into 2D nanosheets.
Foundation of China Youth Science (No. 22205059).


*
ASSOCIATED CONTENT
sı Supporting Information
Notes
The authors declare no competing financial interest.
The Supporting Information is available free of charge at
https://pubs.acs.org/doi/10.1021/acs.cgd.2c00658. ■ ACKNOWLEDGMENTS
Calculation support from the Institute of Theoretical Physics of
Summary of the elastic properties of (TPrA)Cu2(dca)5 the Hubei University of Education is gratefully acknowledged.
and other prototypical 2D structures, material’s informa-
tion of (TPrA)Cu2(dca)5 including crystallographic
information, selected bond lengths and angles, and
asymmetric unit, and characterizations including powder
■ REFERENCES
(1) Cheetham, A. K.; Rao, C. N. R.; Feller, R. K. Structural Diversity
and Chemical Trends in Hybrid Inorganic-Organic Framework
X-ray diffraction, nanoindentation, high-pressure syn- Materials. Chem. Commun. 2006, 4780−4795.
chrotron powder diffraction, SEM, and AFM of the newly (2) Furukawa, H.; Cordova, K. E.; O’Keeffe, M.; Yaghi, O. M. The
synthesized (TPrA)Cu2(dca)5 (PDF) Chemistry and Application of Metal-Organic Frameworks. Science
2013, 341, 974.
Accession Codes (3) Zhou, H. C. J.; Kitagawa, S. Metal-Organic Frameworks (MOFs).
CCDC 2083249 contains the supplementary crystallographic Chem. Soc. Rev. 2014, 43, 5415−5418.
data for this paper. These data can be obtained free of charge via (4) Zhang, J. P.; Zhang, Y. B.; Lin, J. B.; Chen, X. M. Metal Azolate
www.ccdc.cam.ac.uk/data_request/cif, or by emailing data_ Frameworks: From Crystal Engineering to Functional Materials. Chem.
request@ccdc.cam.ac.uk, or by contacting The Cambridge Rev. 2012, 112, 1001−1033.
Crystallographic Data Centre, 12 Union Road, Cambridge (5) Li, W.; Wang, Z. M.; Deschler, F.; Gao, S.; Cheetham, A. K.
CB2 1EZ, UK; fax: +44 1223 336033. Chemically Diverse and Multifunctional Hybrid Organic-Inorganic
Perovskites. Nat. Rev. Mater. 2017, 2, 16099.

■ AUTHOR INFORMATION
Corresponding Authors
(6) Ji, L. J.; Sun, S. J.; Qin, Y.; Li, K.; Li, W. Mechanical Properties of
Hybrid Organic-Inorganic Perovskites. Coord. Chem. Rev. 2019, 391,
15−29.
Wei Li − School of Materials Science and Engineering & Tianjin (7) Bahr, D. F.; Reid, J. A.; Mook, W. M.; Bauer, C. A.; Stumpf, R.;
Key Laboratory of Metal and Molecule-Based Material Skulan, A. J.; Moody, N. R.; Simmons, B. A.; Shindel, M. M.; Allendorf,
M. D. Mechanical Properties of Cubic Zinc Carboxylate IRMOF-1
Chemistry, Nankai University, Tianjin 300350, China; Metal-Organic Framework Crystals. Phys. Rev. B Condens. Matter Mater.
orcid.org/0000-0002-5277-6850; Email: wl276@ Phys. 2007, 76, No. 184106.
nankai.edu.cn (8) Greathouse, J. A.; Allendorf, M. D. Force Field Validation for
Guo-Qiang Feng − Department of Physics and Mechanical & Molecular Dynamics Simulations of IRMOF-1 and Other Isoreticular
Electrical Engineering, Hubei University of Education, Wuhan Zinc Carboxylate Coordination Polymers. J. Phys. Chem. C 2008, 112,
430074, China; Email: gqfeng627@outlook.com 5795−5802.

6992 https://doi.org/10.1021/acs.cgd.2c00658
Cryst. Growth Des. 2022, 22, 6984−6994
Crystal Growth & Design pubs.acs.org/crystal Article

(9) Tan, J. C.; Civalleri, B.; Lin, C. C.; Valenzano, L.; Galvelis, R.; Hydrophobic and Hydrophilic Metal-Organic Frameworks to Form
Chen, P. F.; Bennett, T. D.; Draznieks, C. M.; Wilson, C. M. Z.; Nanosheets. Chem. − Eur. J. 2018, 24, 17986−17996.
Cheetham, A. K. Exceptionally Low Shear Modulus in A Prototypical (26) Saines, P. J.; Steinmann, M.; Tan, J. C.; Yeung, H. H. M.; Li, W.;
Imidazole-Based Metal-Organic Framework. Phys. Rev. Lett. 2012, 108, Barton, P. T.; Cheetham, A. K. Isomer-Directed Structural Diversity
No. 095502. and Its Effect on The Nanosheet Exfoliation and Magnetic Properties of
(10) Chapman, K. W.; Halder, G. J.; Chupas, P. J. Pressure-Induced 2,3-Dimethylsuccinate Hybrid Frameworks. Inorg. Chem. 2012, 51,
Amorphization and Porosity Modification in A Metal-Organic 11198−11209.
Framework. J. Am. Chem. Soc. 2009, 131, 17546−17547. (27) Wang, H. S.; Li, J.; Li, J. Y.; Wang, K.; Ding, Y.; Xia, X. H.
(11) Mishra, M. K.; Ramamurty, U.; Desiraju, G. R. Mechanical Lanthanide-Based Metal-Organic Framework Nanosheets with Unique
Property Design of Molecular Solids. Curr. Opin. Solid State Mater. Sci. Fluorescence Quenching Properties for Two-Color Intracellular
2016, 20, 361−370. Adenosine Imaging in Living Cells. NPG Asia Mater. 2017, 9, No. e354.
(12) Mishra, M. K.; Ramamurty, U.; Desiraju, G. R. Solid Solution (28) Ding, Y.; Chen, Y. P.; Zhang, X.; Chen, L.; Dong, Z.; Jiang, H. L.;
Hardening Molecular Crystals: Tautomeric Polymorphs of Omepra- Xu, H.; Zhou, H. C. Controlled Intercalation and Chemical Exfoliation
zole. J. Am. Chem. Soc. 2015, 137, 1794−1797. of Layered Metal-Organic Frameworks Using A Chemically Labile
(13) Mishra, M. K.; Desiraju, G. R.; Ramamurty, U.; Bond, A. D. Intercalating Agent. J. Am. Chem. Soc. 2017, 139, 9136−9139.
Studying Microstructure in Molecular Crystals With Nanoindentation: (29) Peng, Y.; Li, Y.; Ban, Y.; Yang, W. Two-Dimensional Metal-
Intergrowth Polymorphism in Felodipine. Angew. Chem., Int. Ed. 2014, Organic Framework Nanosheets for Membrane-Based Gas Separation.
53, 13102−13105. Angew. Chem., Int. Ed. 2017, 56, 9757−9761.
(14) Mishra, M. K.; Sanphui, P.; Ramamurty, U.; Desiraju, G. R. (30) Cliffe, M. J.; Castillo-Martinez, E.; Wu, Y.; Lee, J.; Forse, A. C.;
Solubility-Hardness Correlation in Molecular Crystals: Curcumin and Firth, F. C. N.; Moghadam, P. Z.; Fairen-Jimenez, D.; Gaultois, M. W.;
Sulfathiazole Polymorphs. Cryst. Growth Des. 2014, 14, 3054−3061. Hill, J. A.; Magdysyuk, O. V.; Slater, B.; Goodwin, A. L.; Grey, C. P.
(15) Ghosh, S.; Mondal, A.; Kiran, M. S. R. N.; Ramamurty, U.; Metal-Organic Nanosheets Formed via Defect-Mediated Trans-
Reddy, C. M. The Role of Weak Interactions in the Phase Transition formation of A Hafnium Metal-Organic Framework. J. Am. Chem.
and Distinct Mechanical Behavior of Two Structurally Similar Caffeine Soc. 2017, 139, 5397−5404.
Co-crystal Polymorphs Studied by Nanoindentation. Cryst. Growth Des. (31) Wang, X.; Chi, C.; Zhang, K.; Qian, Y.; Gupta, K. M.; Kang, Z.;
2013, 13, 4435−4441. Jiang, J.; Zhao, D. Reversed Thermo-Switchable Molecular Sieving
(16) Raju, K. B.; Ranjan, S.; Vishnu, V. S.; Bhattacharya, M.; Membranes Composed of Two-Dimensional Metal-Organic Nano-
Bhattacharya, B.; Mukhopadhyay, A. K.; Reddy, C. M. Rationalizing sheets for Gas Separation. Nat. Commun. 2017, 8, 14460.
Distinct Mechanical Properties of Three Polymorphs of a Drug Adduct (32) Han, B.; Ou, X.; Deng, Z.; Song, Y.; Tian, C.; Deng, H.; Xu, Y. J.;
by Nanoindentation and Energy Frameworks Analysis: Role of Slip Lin, Z. Ni Metal-Organic Frameworks Monolayers for Photoreduction
Layer Topology and Weak Interactions. Cryst. Growth Des. 2018, 18, of Diluted CO2: Metal Nodes-dependent Activity and Selectivity.
3927−3937. Angew. Chem., Int. Ed. 2018, 57, 16811−16815.
(17) Devarapalli, R.; Kadambi, S. B.; Chen, C. T.; Krishna, G. R.; (33) CrysAlisPro, Version 1.171.39.7a, Oxford Diffraction Ltd.
Kammari, B. R.; Buehler, M. J.; Ramamurty, U.; Reddy, C. M. (34) Sheldrick, G. M. SHELXT-Integrated Space-Group and Crystal-
Remarkably Distinct Mechanical Flexibility in Three Structurally Structure Determination. Acta Crystallogr. A Found. Adv. 2015, 71, 3−8.
Similar Semiconducting Organic Crystals Studied by Nanoindentation (35) Bourhis, L. J.; Dolomanov, O. V.; Gildea, R. J.; Howard, J. A. K.;
and Molecular Dynamics. Chem. Mater. 2019, 31, 1391−1402. Puschmann, H. The Anatomy of A Comprehensive Constrained,
(18) Saha, S.; Mishra, M. K.; Reddy, C. M.; Desiraju, G. R. From Restrained Refinement Program for The Modern Computing Environ-
Molecules to Interactions to Crystal Engineering: Mechanical Proper- ment-Olex2 Dissected. Acta Crystallogr. A Found. Adv. 2015, 71, 59−75.
ties of Organic Solids. Acc. Chem. Res. 2018, 51, 2957−2967. (36) Dolomanov, O. V.; Bourhis, L. J.; Gildea, R. J.; Howard, J. A. K.;
(19) Reddy, C. M.; Krishna, G. R.; Ghosh, S. Mechanical Properties of Puschmann, H. Iotbox. Cif: A Comprehensive CIF Toolbox. J. Appl.
Molecular Crystals-Applications to Crystal Engineering. CrystEng- Crystallogr. 2009, 42, 339−341.
Comm 2010, 12, 2296−2314. (37) Li, W.; Thirumurugan, A.; Barton, P. T.; Lin, Z.; Henke, S.;
(20) Bhattacharya, B.; Roy, D.; Dey, S.; Puthuvakkal, A.; Bhunia, S.; Yeung, H. H. M.; Wharmby, M. T.; Bithell, E. G.; Howard, C. J.;
Mondal, S.; Chowdhury, R.; Bhattacharya, M.; Mandal, M.; Manoj, K.; Cheetham, A. K. Mechanical Tunability via Hydrogen Bonding in
Mandal, P. K.; Reddy, C. M. Mechanical-Bending-Induced Fluo- Metal-Organic Frameworks with The Perovskite Architecture. J. Am.
rescence Enhancement in Plastically Flexible Crystals of a GFP Chem. Soc. 2014, 136, 7801−7804.
Chromophore Analogue. Angew. Chem., Int. Ed. 2020, 59, 19878− (38) Oliver, W. C.; Pharr, G. M. An Improved Technique for
19883. Determining Hardness and Elastic Modulus Using Load and Displace-
(21) Tan, J. C.; Merrill, C. A.; Orton, J. B.; Cheetham, A. K. ment Sensing Indentation. J. Mater. Res. 1992, 7, 1564−1583.
Anisotropic Mechanical Properties of Polymorphic Hybrid Inorganic- (39) Oliver, W. C.; Pharr, G. M. Measurement of Hardness and Elastic
Organic Framework Materials with Different Dimensionalities. Acta Modulus by Instrumented Indentation: Advances in Understanding
Mater. 2009, 57, 3481−3496. And Refinements to Methodology. J. Mater. Res. 2004, 19, 3−20.
(22) Gui, D.; Ji, L.; Muhammad, A.; Li, W.; Cai, W.; Li, Y.; Li, X.; Wu, (40) Feng, G. Q.; Jiang, X. X.; Wei, W. J.; Gong, P. F.; Li, Z. H.; Li, Y.
X.; Lu, P. Jahn-Teller Effect on Framework Flexibility of Hybrid C.; Li, X. D.; Wu, X.; Lin, Z. S.; Li, W.; Lu, P. X. High-Pressure Behavior
Organic-Inorganic Perovskites. J. Phys. Chem. Lett. 2018, 9, 751−755. and Elastic Properties of A Dense Inorganic-Organic Framework.
(23) Lopez-Cabrelles, J.; Manas-Valero, S.; Vitorica-Yrezabal, I. J.; Dalton Trans. 2016, 45, 4303−4308.
Bereciartua, P. J.; Rodriguez-Velamazan, J. A.; Waerenborgh, J. C.; (41) Mao, H. K.; Xu, J.; Bell, P. M. Calibration of The Ruby Pressure
Vieira, B. J. C.; Davidovikj, D.; Steeneken, P. G.; van der Zant, H. S. J.; Gauge to 800 kbar Under Quasi-Hydrostatic Conditions. J. Geophys.
Espallargas, G. M.; Coronado, E. Isoreticular Two-Dimensional Res. 1986, 91, 4673−4676.
Magnetic Coordination Polymers Prepared Through Pre-Synthetic (42) Hammersley, J. Fit2d User Manual; ESRF: Grenoble, France.
Ligand Functionalization. Nat. Chem. 2018, 10, 1001−1007. 1996.
(24) Abherve, A.; Manas-Valero, S.; Clemente-Leon, M.; Coronado, (43) Clark, S. J.; Segall, M. D.; Pickard, C. J.; Hasnip, P. J.; Probert, M.
E. Graphene Related Magnetic Materials: Micromechanical Exfoliation J.; Refson, K.; Payne, M. C. First Principles Method Using CASTEP. Z.
of 2D Layered Magnets Based on Bimetallic Anilate Complexes with Kristallogr. 2005, 220, 567−570.
Inserted [FeIII(acac2-trien)]+ and [FeIII(sal2-trien)]+ molecules. Chem. (44) Payne, M. C.; Teter, M. P.; Allen, D. C.; Arias, T. A.;
Sci. 2015, 6, 4665−4673. Joannopoulos, J. D. Iterative Minimization Techniques for ab Initio
(25) Ashworth, D. J.; Cooper, A.; Trueman, M.; Al-Saedi, R. W. M.; Total-Energy Calculations: Molecular Dynamics And Conjugate
Smith, L. D.; Meijer, A. J. H. M.; Foster, J. A. Ultrasonic Exfoliation of Gradients. Rev. Mod. Phys. 1992, 64, 1045.

6993 https://doi.org/10.1021/acs.cgd.2c00658
Cryst. Growth Des. 2022, 22, 6984−6994
Crystal Growth & Design pubs.acs.org/crystal Article

(45) Perdew, J. P.; Wang, Y. Pair-Distribution Function And Its (66) Gao, H.; Li, C.; Li, L.; Wei, W.; Tan, Y.; Tang, Y. High Pressure
Coupling-Constant Average For The Spin-Polarized Electron Gas. And Elastic Properties of A Guanidinium-Formate Hybrid Perovskite.
Phys. Rev. B: Condens. Matter Mater. Phys. 1992, 46, 12947. Dalton Trans. 2020, 49, 7228−7233.
(46) Pickard, C. J.; Needs, R. J. Aluminium at Terapascal Pressure. (67) Guo, T. M.; Gao, F. F.; Li, Z. G.; Liu, Y.; Yu, M. H.; Li, W.
Nat. Mater. 2010, 9, 624−627. Mechanical and Acoustic Properties of A Hybrid Organic-Inorganic
(47) Monkhorst, H. J.; Pack, J. D. Special Point for Brillouin-Zone Perovskite, TMCM-CdCl3 With Large Piezoelectricity. APL Mater.
Integrations. Phys. Rev. B 1976, 13, 5188−5192. 2020, 8, 101106.
(48) Marmier, A.; Lethbridge, Z. A. D.; Walton, R. L.; Smith, C. W.; (68) Blakslee, O. L.; Proctor, D. G.; Seldin, E. J.; Spence, G. B.; Wang,
Parker, S. C.; Evans, K. E. ELAM: A Computer Program for The T. Elastic Constants of Compression-Annealed Pyrolytic Graphite. J.
Analysis And Representation of Anisotropic Elastic Properties. Comput. Appl. Phys. 1970, 41, 3373−3382.
Phys. Commun. 2010, 181, 2102−2115. (69) Wei, Q.; Peng, X. Superior Mechanical Flexibility of
(49) Ramamurty, U.; Jiang, J. Nanoindentation for Probing The Phosphorene And Few-Layer Black Phosphorus. APL Mater. 2014,
Mechanical Behavior of Molecular Crystals-A Review of The 104, 251915.
Technique And How To Use It. CrystEngComm 2014, 16, 12−23. (70) Feldman, J. L. Elastic Constants of 2H-MoS2 And 2H-NbSe2
(50) Varughese, S.; Kiran, M. S. R. N.; Ramamurty, U.; Desiraju, G. R. Extracted From Measured Dispersion Curves And Linear Compres-
Nanoindentation In Crystal Engineering: Quantifying Mechanical siblities. J. Phys. Chem. Solids 1976, 37, 1141−1144.
Properties Of Molecular Crystals. Angew. Chem., Int. Ed. 2013, 52, (71) Feng, G.; Qin, Y.; Ran, C.; Ji, L.; Dong, L.; Li, W. Structural
2701−2712. Evolution And Photoluminescence Properties of A 2D Hybrid
(51) Varughese, S.; Kiran, M. S. R. N.; Solanko, K. A.; Bond, A. D.; Perovskite Under Pressure. APL Mater. 2018, 6, 114201.
Ramamurty, U.; Desiraju, G. R. Interaction Anisotropy And Shear (72) Pugh, S. F. XCII. Relations Between The Elastic Moduli And The
Instability Of Aspirin Polymorphs Established by Nanoindentation. Plastic Properties of Polycrystalline Pure Metals. Philos. Mag. J. Sci.
Chem. Sci. 2011, 2, 2236−2242. 2009, 45, 823−843.
(52) Kiran, M. S. R. N.; Varughese, S.; Ramamurty, U.; Desiraju, G. R.
Effect Of Dehydration On The Mechanical Properties Of Sodium
Saccharin Dihydrate Probed With Nanoindentation. CrystEngComm
2012, 14, 2489.
(53) Tan, J. C.; Cheetham, A. K. Mechanical Properties of Hybrid
Inorganic-Organic Framework Materials: Establishing Fundamental
Structure-Property Relationships. Chem. Soc. Rev. 2011, 40, 1059.
(54) Li, W.; Henke, S.; Cheetham, A. K. Research Update: Mechanical
Properties Of Metal-Organic Frameworks-Influence Of Structure And
Chemical Bonding. APL Mater. 2014, 2, 123902.
(55) Ji, L. J.; Qin, Y.; Gui, D.; Li, W.; Li, Y. C.; Li, X. D.; Lu, P. X.
Quantifying The Exfoliation Ease Level Of 2D Materials Via
Mechanical Anisotropy. Chem. Mater. 2018, 30, 8732−8738.
(56) Qin, Y.; Lv, Z. X.; Chen, S. G.; Li, W.; Wu, X.; Ye, L.; Li, N.; Lu, P.
X. Tuning Pressure-Induced Phase Transitions, Amorphization, and Recommended by ACS
Excitonic Emissions of 2D Hybrid Perovskites via Varying Organic
Amine Cations. J. Phys. Chem. C 2019, 123, 22491−22498. Biaxial Negative Thermal Expansion in an Organic
(57) Cliffe, M. J.; Goodwin, A. L. PASCal: A Principal Axis Strain Imidazolium Salt: Modulation of the Thermal Expansion
Calculator For Thermal Expansion And Compressibility Determi- Property by Methyl Substitution
nation. J. Appl. Crystallogr. 2012, 45, 1321−1329.
(58) Li, Z. G.; Li, K.; Dong, L. Y.; Guo, T. M.; Azeem, M.; Li, W.; Bu, Stanzin Chuskit, Dinabandhu Das, et al.
X. H. Acoustic Properties Of Metal-Organic Frameworks. Research FEBRUARY 20, 2023
CRYSTAL GROWTH & DESIGN READ
2021, 2021, No. 9850151.
(59) Kai, L.; Dong, L.; Xu, H.; Qin, Y.; Li, Z.; Azeem, M.; Li, W.; Bu, X.
Electronic Structures And Elastic Properties Of A Family Of Metal-Free Piezoelectric Response of Plastic Ionic Molecular Crystals:
Perovskites. Mater. Chem. Front. 2019, 3, 1678. Role of Molecular Rotation
(60) Ortiz, A. U.; Boutin, A.; Fuchs, A. H.; Coudert, F. X. Anisotropic Elin D. Sødahl, Kristian Berland, et al.
Elastic Properties Of Flexible Metal-Organic Frameworks: How Soft JANUARY 06, 2023
And Soft Porous Crystals. Phys. Rev. Lett. 2012, 109, No. 195502. CRYSTAL GROWTH & DESIGN READ
(61) Qin, Y.; Li, Z. G.; Gao, F. F.; Chen, H.; Li, X.; Xu, B.; Li, Q.; Jiang,
X.; Li, W.; Wu, X.; Quan, Z.; Ye, L.; Zhang, Y.; Lin, Z.; Pedesseau, L.; H/F-Substitution-Induced Confinement Effect for Designing
Even, J.; Lu, P.; Bu, X. H. Dangling Octahedra Enable Edge States in 2D High-Phase-Transition-Temperature Organic–Inorganic
Lead Halide Perovskites. Adv. Mater. 2022, 34, No. 2201666. Hybrid Materials
(62) Li, K.; Qin, Y.; Li, Z. G.; Guo, T. M.; An, L. C.; Li, W.; Li, N.; Bu,
X. H. Elastic Properties Related Energy Conversions of Coordination Yan-Ning Wang, Li-Zhuang Chen, et al.
Polymers And Metal-Organic Frameworks. Coord. Chem. Rev. 2022, SEPTEMBER 15, 2022
CRYSTAL GROWTH & DESIGN READ
470, No. 214692.
(63) An, L. C.; Li, K.; Li, Z. G.; Zhu, S. L.; Li, Q. T.; Zhang, Z. Z.; Ji, L.
J.; Li, W.; Bu, X. H. Engineering Elastic Properties Of Isostructural Construction of Cu(I)-Organic Frameworks with Effective
Molecular Perovskite Ferroelectrics via B-Site Substitution. Small 2021, Sorption Behavior for Iodine and Congo Red
17, No. e2006021. Zhen-Zhen Xue, Jie Pan, et al.
(64) Errandonea, D.; Meng, Y.; Somayazulu, M.; Hausermann, D. AUGUST 23, 2022
Pressure-Induced α⃗ ω Transition in Titanium Metal: A Systematic INORGANIC CHEMISTRY READ
Study of The Effects of Uniaxial Stress. Phys. B 2005, 355, 116−125.
(65) Shen, Y. R.; Kumar, R. S.; Pravica, M.; Nicol, M. F.
Get More Suggestions >
Characteristics of Silicone Fluid As A Pressure Transmitting Medium
in Diamond Anvil Cells. Rev. Sci. Instrum. 2004, 75, 4450−4454.

6994 https://doi.org/10.1021/acs.cgd.2c00658
Cryst. Growth Des. 2022, 22, 6984−6994

You might also like