Next Article in Journal
Effects of Sputtering Parameters on AlN Film Growth on Flexible Hastelloy Tapes by Two-Step Deposition Technique
Previous Article in Journal
Metals Recovery from Artificial Ore in Case of Printed Circuit Boards, Using Plasmatron Plasma Reactor
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Nitrogen-Rich Energetic Metal-Organic Framework: Synthesis, Structure, Properties, and Thermal Behaviors of Pb(II) Complex Based on N,N-Bis(1H-tetrazole-5-yl)-Amine

1
Research Center of Laser Fusion, China Academy of Engineering Physics, Mianyang 621010, China
2
State Key Laboratory Cultivation Base for Nonmetal Composites and Functional Materials, Southwest University of Science and Technology, Mianyang 621010, China
3
School of National Defence Science and Technology, Southwest University of Science and Technology, Mianyang 621010, China
4
Institute of Chemical Materials, Chinese Academy of Engineering Physics, Mianyang 621010, China
*
Authors to whom correspondence should be addressed.
Materials 2016, 9(8), 681; https://doi.org/10.3390/ma9080681
Submission received: 20 June 2016 / Revised: 21 July 2016 / Accepted: 25 July 2016 / Published: 10 August 2016

Abstract

:
The focus of energetic materials is on searching for a high-energy, high-density, insensitive material. Previous investigations have shown that 3D energetic metal–organic frameworks (E-MOFs) have great potential and advantages in this field. A nitrogen-rich E-MOF, Pb(bta)·2H2O [N% = 31.98%, H2bta = N,N-Bis(1H-tetrazole-5-yl)-amine], was prepared through a one-step hydrothermal reaction in this study. Its crystal structure was determined through single-crystal X-ray diffraction, Fourier transform infrared spectroscopy, and elemental analysis. The complex has high heat denotation (16.142 kJ·cm−3), high density (3.250 g·cm−3), and good thermostability (Tdec = 614.9 K, 5 K·min−1). The detonation pressure and velocity obtained through theoretical calculations were 43.47 GPa and 8.963 km·s−1, respectively. The sensitivity test showed that the complex is an impact-insensitive material (IS > 40 J). The thermal decomposition process and kinetic parameters of the complex were also investigated through thermogravimetry and differential scanning calorimetry. Non-isothermal kinetic parameters were calculated through the methods of Kissinger and Ozawa-Doyle. Results highlighted the nitrogen-rich MOF as a potential energetic material.

1. Introduction

In the past decade, metal-organic frameworks (MOFs) have elicited much interest in chemistry, material science, medicine, and environmental science [1,2,3,4,5,6,7,8,9] because of their stable architectures, controllable structures, modifiable properties, and potential applications in gas storage [10,11,12,13], chemical separation [14,15,16,17], catalysis [18,19,20,21], chemical sensor technology [5], drug delivery [22,23,24], and so on. Many investigators have recently demonstrated the possibility of using nitrogen-rich MOFs as high explosives [25,26,27,28,29,30,31,32,33,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48]. MOFs consist of metal ions (Pb2+, Ag+, etc.), energetic anions (e.g., N3 and NO3), or simple energetic ligands (triazole, tetrazole, tetrazine, hydrazine, etc.). The network structures of energetic MOFs (E-MOFs) can be designed as 1D, 2D, or 3D architectures (Figure 1) depending on the metal ion geometry and binding mode of the bridging energetic ligands.
Several 1D E-MOFs (CHP: CoN10H20Cl2O8, N% = 33.49%, ρ = 1.948 g·cm−3, NHP: NiN10H20O8Cl2, NHN: NiN8H12O6; Figure 2) and 2D E-MOFs (CHHP: Co4C4H48N24O26Cl4, N% = 27.41%, ρ = 2.0 g·cm−3; ZnHHP: Zn4C4H40N20O26Cl4, N% = 23.57%, ρ = 2.117 g·cm−3) were developed by Hope-Weeks et al. through a combination of metal ion, energetic anion (ClO4 and NO3), and non-bridged ligand (hydrazine) [25,27]. 1D E-MOFs have high detonation heat but are very sensitive. The sensitivity and detonation heat of 2D E-MOFs are lower than those of 1D E-MOFs. Different from 1D linear and 2D layered structures, 3D frameworks possess more complicated connection modes, which could further enhance structural reinforcement and thus improve the stability and energetic properties [29,49]. In 2013, two 3D E-MOFs, namely, CuC12H12N26O6 (N% = 53.35%, ρ = 1.68 g·cm−3, IS = 22.5 J) and AgC6H6N13O3 (N% = 53.35%, ρ = 2.16 g·cm−3, IS = 30 J), were introduced by Pang et al. for the first time [28]. The sensitivities of these 3D MOFs are significantly lower than those of reported energetic coordination polymers, such as 1D (CHP, IS = 0.5 J) and 2D (ZnHHP, IS = 2.5 J; CHHP, IS = 0.8 J) MOFs. An increasing number of investigations on E-MOFs as new-generation high explosives were reported by Chen et al. [32,33,36,38,39,40,42,44,45,50,51], Pang et al. [28,48,52], Shreeve et al. [49,53], and so on [41,47,54,55,56] because of the advantages of 3D MOFs.
N,N-Bis(1H-tetrazole-5-yl)-amine (H2bta), a compound with high nitrogen content (N% = 82.34%), had its first single crystal structure and copper complexes reported by Klapötke et al. [57,58]. Shreeve et al. conducted extensive research on the energetic salts of H2bta [59,60,61] and found that the salts exhibit excellent energetic properties. The investigations also showed that H2bta may be an excellent energetic ligand to construct 3D MOFs for the following reasons [51]. First, the rigid structure of H2bta cannot only promote structural stability but can also improve the energetic performance of MOFs. Second, the versatile chelating-bridging coordination modes are propitious to the construction of high-dimensional 3D MOFs. Third, nitrogen atoms are involved in hydrogen-bond motifs to capture energetic moieties. Fourth, the predominant decomposition products are environmentally benign nitrogen gases. Fifth, the azide group hidden in aminotetrazole is the best stable moiety by virtue of the extended 6π system. Therefore, many MOFs based on H2bta, such as Cu(II) [39,62,63,64], Co(II,III) [36,50], Fe(III) [65], Mn(II) [66,67], Zn(II) [46,66,68,69], Cd(II) [66] MOFs, and others [70], have been obtained and investigated.
Pb(II)-bta MOFs (The Cambridge Crystallographic Data Centre (CCDC) numbers 650737, 721842, 721843) were synthesized originally via a two-step approach, and there were no reports about their energetic properties [71]. In this study, [Pb(bta)·2H2O]n (CCDC number 1478651) was synthesized via a one-step hydrothermal reaction without any other assistant ligands (such as 2,2-bipyridine, 1,10-phenanthroline). The single crystal X-ray experiment revealedthe coordination mode of bta2− with Pb2+. The energetic properties (detonation velocity, detonation pressure, and impact sensitivity) and thermal behavior of [Pb(bta)·2H2O]n were also investigated. The thermodynamic parameters were obtained based on the reaction thermodynamic and kinetic equations. As expected, the complex exhibited high thermostability, excellent detonation properties, and acceptable sensitivity to impact. These featuressuggest potential applications as an energetic material.

2. Results and Discussion

General caution: H2bta and its derivatives are potentially explosive and should be handled in small quantities. Appropriate safety precautions should be taken, and larger scale synthesis is not recommended.

2.1. Synthesis of the Complex

H2bta·H2O was synthesized according to the literature [58]. Herein, three methods were used to synthesize Pb(II) coordination compounds based on H2bta, but only one route obtained the target complex successfully. As shown in Scheme 1, a mixture of Pb(NO3)2 (0.1 mmol) and H2bta·H2O (0.13 mmol) in H2O (4 mL) was sealed in a 10-mL Teflon-lined stainless autoclave and heated at 130 °C under autogenous pressure for three days and then cooled to room temperature over a further threedays. Colorless prismatic single-crystals suitable for X-ray diffraction were obtained.

2.2. Crystal Structure of the Complex

The X-ray crystallographic and structural refinement data are summarized in Table 1, and the structures are shown in Figure 3, Figure 4 and Figure 5. Further information on thecrystalstructure determination is providedin the Supplementary Materials (Tables S1–S6). Analysis of the X-ray crystallographic data for the complex shows that it crystallizes in the monoclinic space group P21/n with a calculated density of 3.250 g·cm−3 based on four molecules packed in the unit-cell volume of 806.0(10) Å3. Density is a highly important physical property of energetic materials. Herein, because metal coordination can improve the densities of energetic materials, the density values of complex is much higher than that of the free ligand (1.693 g·cm−3).
The asymmetric unit is crystallographically independent with one Pb(II) ion, one bta2− ligand, and two coordination water molecules (Figure 3a). Figure 3a shows that each Pb(II) ion is coordinated by three nitrogen atoms from two bta2− ligands and three oxygen atoms from water molecules. Anirregular octahedral geometry is exhibited. Compared with the structure of H2bta [58], all of the bond lengths and angles are slightly changed, which may be caused by the negative charge on the anion rings and thecoordination environment. The C–N bond lengths of C1–N1 [1.322(9) Å], C1–N4 [1.322(9) Å], C1–N5 [1.379(9) Å], C2–N5 [1.377(8) Å], C2–N6 [1.330(8) Å], and C2–N9 [1.328(8) Å] are between the standard C–N single bond (1.47 Å) and standard C=N double bond (1.32 Å) lengths and are indicative of an aromatic system [72,73]. Meanwhile, the N–N bond lengths of N1–N2 [1.376(7) Å], N2–N3 [1.284(9) Å], N3–N4 [1.350(8) Å], N6–N7 [1.347(9) Å], N7–N8 [1.306(8) Å], and N8–N9 [1.369(8) Å] also fit between the standard N–N single bond (1.45 Å) and standard N=N double bond (1.25 Å) lengths [74,75], which further confirmstheconjugated and aromatic system. In addition, the bond angles C2–N5–C1 [125.0(6)°], N4–C1–N5 [112.8(6)°], N5–C2–N9 [127.2(6)°], and N6–C2–N9 [111.9(6)°] are slightly larger than thoseof H2bta because of the influence of the coordination environment. Figure 3b shows that the coordination mode of the ligand has three coordinated nitrogen atoms (N1, N9, and N6) in each bta2−. Atoms N1 and N9 in the ligand adopt chelating modes to connect to Pb(II) ion, whereas atom N6 adopts monodentate bridging modes to link with other Pb(II) ions.
The packing diagram of the complex viewed down the a-axis, b-axis, and the a–c diagonal isshown in Figure 4. The adjacent Pb(II) ions are bridged by two oxygen atoms from water in an antiparallel manner, with a Pb...Pb separation distance of 4.484 Å, Pb1–O1W distance of 2.670 Å, Pb1–O2W distance of 2.586 Å, O2W–Pb1–O2W angle of 69.01°, and O2W–Pb1–O1W angle of 144.05°. The view down the b-axis shows that a series of parallel rhombiare formed by the adjacent Pb1 and O2w. The torsion angles [i.e., N3–N2–N1–C1 (−0.2°), N9–Pb1–N1–N2 (179.2°), N1–N2–N3–N4 (−0.7°), and C2–N5–C1–N4 (179.3°)] are close to ±180° and 0°, which illustrates that Pb(II) and its chelating mode ligand are strictly coplanar. In addition, five types of hydrogen bonds [N5-H5A...N83# = 3.032 Å, O1W-H1WA...N24# = 2.841 Å, O1W-H1WB...N35# = 3.411 Å, O2W-H2WA...N36# = 2.856 Å, and O2W-H2WB...O1W7# = 2.715 Å; symmetry transformations used to generate equivalent atoms: (3# x − 1/2, −y + 3/2, z − 1/2),(4# −x + 1/2, y − 1/2, −z + 3/2), (5# −x, −y + 2, −z + 1), (6# −x − 1/2, −y − 1/2, −z + 3/2), and (7# x − 1, y, z)] (others are listed in the Supplementary Materials) exist in the target complex and further enhance the structural reinforcement. The complex [Pb(bta)·2H2O]n has a good symmetrical and ordered 3D energetic framework.
The coordination polyhedron geometry of the complex is shown in Figure 5. Pb(bta)·2H2O is a six-coordinate complex, and its polyhedron with Pb(II) as the center is an irregular octahedral geometry. The basal plane of the octahedron is formed by three nitrogen atoms and O2’ atoms, and the vertexes are occupied by O1 and O2 atoms (Figure 5a). Two adjacent octahedrons share the O2...O2’ edge. Figure 5c shows that the planes that cross the Pb(II) ions in two interval polyhedrons are parallel. The two planes through the Pb(II) ions in two adjacent polyhedrons intersect, and the dihedral angle of the two planes is 33.8°.

2.3. Thermal Decomposition and Non-Isothermal Kinetics Analysis

2.3.1. Thermal Decomposition

Approximately 1.0 mg of the complex was tested through differential scanning calorimetry (DSC) in an open crucible at a heating rate of 5 K·min−1 under nitrogen atmosphere to determine the melting points and decomposition temperatures. One endothermic process (loss of crystal water) peak and one exothermic process (decomposition) peak at 412.4 K and 614.9 K, respectively, are visible in the DSC curve (Figure 6). The relevant exothermic enthalpy change of the compound is 352.1 kJ·mol−1. Therefore, this metal–organic crystal containing the bta2− ligand possesses sufficient thermal stability to be an energetic material. In addition, the effect on Tp by the particle size were also investigated (Figures S2 and S3). The thermal behavior of millimeter sized single crystals and micron sized crystals were studied by DSC at the heating rate of 10 K·min−1 under nitrogen atmosphere. The peak decomposition temperature of micron sized crystals MOF (about 10 μm) is 615.4 K, which is lower than that of single crystals MOF (Tp = 620.1 K). The results are consistent with the reference by Cacho-Bailo et al. [76].
The decomposition process of the complex was also investigated through the rmogravimetric analysis (TGA) at a heating rate of 5 K·min−1 in nitrogen atmosphere. Figure 7 shows that the decomposition process of the complex can be divided into two steps, and the total mass loss is 48.09%. The first process in the range of 393 K–443 K was confirmed as the loss of crystal water (observed 9.24%, calculated 9.13%). This result indicates that the water molecules in the title complex were stable before 443 K. The second process from 581 to 657 K was considered the collapse of the structures related to the nitrogen-rich bta2− groups.

2.3.2. Non-Isothermal Kinetics Analysis

Kissinger’s [77] and Ozawa’s methods [78,79] were used to determine the kinetics parameters based on the exothermic peaks temperature measured from DSC curves with four different heating rates (5, 10, 15, 20 K·min−1; Figure 8).
Using the values of peak temperature, Kissinger Equation (Equation (1) in Section 3) and Ozawa-Doyle (Equation (2) in Section 3), the apparent activation energy (E) and pre-exponential factor (A) were calculated. The calculated results including linear correlation coefficient r are shown in Table 2. The results show that the value of E (about 430 kJ·mol−1) is higher than that of 1,3,5-trinitrohexahydro-1,3,5-triazine (RDX, about 142 kJ·mol−1) [62], cyclotetramethylene tetranitramine (HMX, about 238 kJ·mol−1) [62] and 2,4,6,8,10,12-hexanitro-2,4,6,8,10,12-hexaazatetracyclo [5.5.0.0.0] dodecane (CL-20, about 200 kJ·mol−1) [80,81], which is in accordance with its excellent thermostability.

2.4. Energetic Properties

2.4.1. Heat Detonation of Complex

We selected anidentical method for CHP and NHPto estimate the detonationheat (∆Hdet) of the compound and compare it withthevalues for E-MOFs and classical energetic materials [25]. For the complex, Pb, N2, H2O, NH3, and C were assumed to be the final decomposition products of the organic part of the framework, and all non-metal-containing products, including water, were regardedas gas. The detonation reaction considered for the compound is described by Equation (3) (see Section 3), and the heat detonation value was obtained with Equation (4) (see Section 3), which was developed from known ∆Hdet data for 11 commonly used high explosives.
Heat of detonation (∆Hdet) was calculated to be 4.966 kJ·g−1. It is higher than those of MOFs ([Pb(Htztr)(O)]n, 0.94 kJ·g−1; CHHP, 3.14 kJ·g−1; ZnHHP, 2.93 kJ·g−1) but lower than those of RDX (5.799 kJ·g−1) and HMX (5.523 kJ·g−1) (see in Table 3). Figure 9 shows that despite the lack of advantage of the ∆Hdet of the per gram complex, the ∆Hdet of per cm3 (16.142 kJ·cm−3) is higher than that of traditional explosives and E-MOFs, except for ATRZ-1 (ATRZ: 4,4’-azo-1,2,4-triazole) because of its high density.

2.4.2. Detonation Properties and Sensitivity

The performance of a high explosiveischaracterized by its detonation velocity, D (km·s−1), and detonation pressure, P (GPa). The D and P of the complex were calculated with the Kamlet–Jacobs equations (see Equations (5)–(7) in Section 3), which are usually applied to E-MOFs reported previously. Table 3 shows a comparison of the physicochemical properties of several energetic materials and the complex. The D and P of the complex are 8.963 km·s−1 and 47.47 GPa, respectively. Its D is higher than that of HMX (8.900 km·s−1), RDX (8.600 km·s−1), and other E-MOFs (6.205–8.226 km·s−1), except for ATRZ-1 (9.160 km·s−1).
Sensitivity deserves significant attention from researchers because it is closely linked with the safety of handling and applying explosives. The impact sensitivity (IS) of the compound was investigated for initial safety testing. Table 3 provides a summary of the data collected. The IS of [Pb(bta)·2H2O]n is more than 40 J, whereas the IS of RDX is 7.4 J under the same test condition. The IS of the complex is more insensitive than traditional explosives (HMX, 7.4 J; RDX, 7.4J) and reported energetic coordination polymers, such as 1D (CHP, IS = 0.5 J), 2D (ZnHHP, IS = 2.5 J; CHHP, IS = 0.8 J), and 3D (ATRZ-1, IS = 22.5 J) MOFs. These results reveal that the compound is insensitive to external stimuli because of the stabilized 3D covalent framework, in which the molecules are more rigid than those in 1D or 2D structures. Thus, [Pb(bta)·2H2O]n can be classified as an impact-insensitive energetic material.

3. Materials and Methods

The FT-IR spectrum was recorded on Nicolet 380 FT-IR spectrophotometer (Thermo Fisher Nicolet, Waltham, MI, USA) employing a KBr matrix with a resolution of 4 cm−1, in the wavelength range of 400 cm−1 to 4000 cm−1. Elemental analysis was performed on a Vario ELCUBE Elemental Analyzer (Elementar, Hanau, Germany). DSC was performed by a Q200 DSC instrument (TA Instruments, New Castle, DE, USA) at a heating rate of 5, 10, 15 and 20 K·min−1, respectively, in flowing high-purity nitrogen. Approx. 1.0 mg sample was sealed in aluminum pans in the temperature range of 313 to 773 K for DSC experiments. The sensitivity to impact stimuli was determined by fall hammer apparatus according tothe standard staircase method (GJB of China). With a step of 0.04 cm, 50 ± 1 mg of test specimens were used and a 10-kg drop weight was allowed to fall freely from different heights. The results were reported in terms of height for 50% probability of explosion (h50%). The picture of micron sized crystal MOF was obtained by a microscope (×1000). Molecular structure packing diagram and coordination polyhedron geometry of [Pb(bta)·2H2O]n were drawn by Mercury and Diamond software (Diamond 3.1, Crystal Impact GbR, Bonn, Germany).

3.1. Synthesis of [Pb(bta)·2H2O]n

Method 1: a mixture of Pb(NO3)2 (33 mg, 0.1 mmol) and H2bta·H2O (22 mg, 0.13 mmol) in H2O (4 mL) was sealed in a 10-mL Teflon-lined stainless autoclave and heated at 200 °C under autogenous pressure for 3 days and then cooled to room temperature over a further 3 days. Two colorless single-crystals (needle and prismatic crystals) were picked up and a single crystal test indicated they are two different crystals of Pb(N3)2. See the structures in Supplementary Materials (Figure S1, Tables S8–S12).
Method 2: a suspension of H2O (8 mL) and Pb(II) complex (30 mg), which was obtained from Na2bta and Pb(NO3)2 as raw materials by means of metathesis reaction, was sealed in a 10 mL Teflon-lined stainless autoclave. Both experiments were heated at 130 °C and 200 °C, respectively, under autogenous pressure for 3 days and then cooled to room temperature over a further 3 days. Unfortunately, there were powder instead of crystals in Teflon autoclave.
Method 3: a mixture of Pb(NO3)2 (33 mg, 0.1 mmol) and H2bta·H2O (22 mg, 0.13 mmol) in H2O (4 mL) was sealed in a 10-mL Teflon-lined stainless autoclave and heated at 130 °C under autogenous pressure for 3 days and then cooled to room temperature over a further 3 days. Colorless prismatic single-crystals were picked up and yield 23.8 mg (61% based on Pb). DSC (5 K·min−1): 614.9 K(dec.). FT-IR (KBr) : 3432(vs), 3292(s), 2921(m), 2854(m), 1623(vs), 1525(m), 1500(m), 1420(m), 1314(m), 1230(w), 1127(w), 1115(w), 1069(w), 1015(w), 794(m), 740(m). Elemental analysis (C2H5N9O2Pb, 394.34) Calcd: C 6.09%, H 1.28%, N 31.98%; Found: C 6.11%, H 1.30%, N 31.89%.

3.2. Single-Crystal X-ray Diffraction Analyses

The single-crystal X-ray experiments were performed on a Smart Apex CCD diffractometer (Bruker) (Bruker, Karlsruhe, Germany) equipped with graphite monochromatized Mo radiation (λ = 0.71073 A) using the ω and ϕ scan mode. The structure was solved by direct methods using SHELXS-97 (Göttingen, Germany) [82] and refined by means of full-matrix least-squares procedures on F2 with the SHELXL-97 program [83]. All non-H atoms were located using subsequent Fourier-difference methods and refined anisotropically. In all cases, hydrogen atoms were placed in their calculated positions and thereafter allowed to ride on their parent atoms.

3.3. Equations for Calculating Non-Isothermal Kinetics

The Kissinger (1) and Ozawa-Doyle (2) Equations are as follows:
ln [ β / T p 2 ] = ln [ A R / E a ] - [ E a / R T p ] ,
lg β = C - 0 . 4567 E a / R T ,
where β is the heating rate; Tp is the peak temperature; A is the pre-exponential factor; E is the apparent activation energy; and R is the gas constant (8.314 J·K−1·mol−1). Linear relationship of ln(β/Tp2) and lg (β) vs. 1/Tp are shown in Supplementary Materials (Figure S4).

3.4. Calculation for Heat of Detonation

The complete detonation reactions are described by Equation (3). According to Ref. [25], Density Functional Theory (DFT) was used to calculate the energy of detonation (ΔEdet), from which ΔHdet was estimated by using a linear correlation Equation (4), and the calculated parameters was listed in Table S7. The DFT calculation for [Pb(bta)·2H2O]n was performed with the code DMOl3 [84] under 3D periodic boundary conditions employing the Monkhorst–Pack multiple K-point sampling of the Brillouin zone [85] and the Perdew–Becke–Ezerhoff (PBE) exchange-correlation function [86]:
C 2 H 5 N 9 O 2 Pb Pb + 13 3 N 2 + 2 H 2 O + 1 3 N H 3 + 2 C ,
Δ H det = 1 . 127   Δ E det +   0.046 .

3.5. Calculation for Detonation Velocity and Detonation Pressure

The D and P of the complex were calculated by Kamlet–Jacbos Equations [28] as follows, which were usually applied to the E-MOFs reported previously:
D = 1 . 01 Φ 1 / 2 ( 1 + 1 . 30 ρ ) ,
P = 1 . 558 Φ ρ 2 ,
Φ = 31 . 68 N ( M Q ) 1 / 2 .
where D is detonation velocity (km·s−1); P is detonation pressure (GPa); N is moles of detonation gases per gram of explosive; M is average molecular weight of the gases; Q is heat of detonation (kcal·g−1); ρ is density of explosive (g·cm−3); and Φ is a parameter determined by N, M and Q.

4. Conclusions

We successfully synthesized a nitrogen-rich E-MOF, namely, Pb(bta)·2H2O [N% = 31.98%, H2bta = N,N-Bis(1H-tetrazole-5-yl)-amine]. It was characterized through various techniques, such as elemental analyses, Fourier transform infrared spectroscopy, TG, DSC, and single crystal X-ray diffraction. X-ray single crystal structure analysis showed that the crystal of the complex in the monoclinic space group P21/n has a calculated density of 3.250 g·cm−3. The DSC curve indicated that it has good thermostability. One endothermic process (around 412.4 K) and one exothermic process (around 614.9 K) exist at the heating rate 5 K·min−1. The calculated results showed that the detonation heat, detonation pressure, and velocity are 4.966 kJ·g−1 (16.142 kJ·cm−3), 43.47 GPa, and 8.963 km·s−1, respectively. The sensitivity test showed that the complex is an impact-insensitive material (IS > 40 J). The thermal decomposition process and kinetic parameters of the complex were also investigated through TG and DSC. The non-isothermal kinetic parameters were calculated withthe methods of Kissinger and Ozawa-Doyle. The activation energy value (about 430 kJ·mol−1) is higher than that of RDX, HMX, and CL-20. Excellent impact sensitivity and high thermal stability depend on good detonation properties. The 3D MOF in this study has potential applications as an energetic material.

Supplementary Materials

The following are available online at www.mdpi.com/1996-1944/9/8/681/s1. Figure S1: Ball-and-stick molecular structure and packing diagram of Pb(N3)2, Figure S2: The pictures of single crystal MOF and micron sized crystal MOF, Figure S3: DSC curves of single crystal MOFand micron sized crystal MOFat the heating rate of 10 K·min−1, Figure S4: Linear relationship of ln /Tp2) and lg (β) vs. 1/Tp, Table S1: Atomic coordinates and equivalent isotropic displacement parameters for [Pb(bta)·2H2O]n, Table S2: Bond lengths and angles for [Pb(bta)·2H2O]n, Table S3: Anisotropic displacement parameters for [Pb(bta)·2H2O]n, Table S4: Hydrogen coordinates and isotropic displacement parameters for [Pb(bta)·2H2O]n, Table S5: Torsion angles for [Pb(bta)·2H2O]n, Table S6: Hydrogen bonds for [Pb(bta)·2H2O]n, Table S7: Calculated parameters used in the detonation reactions, Table S8: Crystal data and structure refinement for Pb(N3)2, Table S9: Atomic coordinates and equivalent isotropic displacement parameters for Pb(N3)2, Table S10: Bond lengthsand anglesfor Pb(N3)2, Table S11: Anisotropic displacement parameters for Pb(N3)2, Table S12: Torsion angles for Pb(N3)2.

Acknowledgments

We are grateful for financial support from the National Natural Science Foundation of China (project No. 51372211), the Open Project of State Key Laboratory Cultivation Base for Nonmetal Composites and Functional Materials (project No. 14zdfk05), Southwest University of Science and Technology Outstanding Youth Fundation (project No. 13zx9107), and the Institute of Chemical Materials, China Academy of Engineering Physics.

Author Contributions

Qiangqiang Liu and Bo Jin conceived and designed the experiments; Qiangqiang Liu and Yu Shang performed the experiments; Qiangqiang Liu, Qingchun Zhang, Yu Shang, Bo Jin and Rufang Peng analyzed the data; Zhicheng Guo and Bisheng Tan calculated the Denotation Heat of the complex and analyzed those data. Bo Jin and Rufang Peng directed the project and contributed reagents/materials/analysis tools; Qiangqiang Liu and Bo Jin wrote the paper.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Eddaoudi, M.; Moler, D.B.; Li, H.; Chen, B.; Reineke, T.M.; O’keeffe, M.; Yaghi, O.M. Modular chemistry: Secondary building units as a basis for the design of highly porous and robust metal-organic carboxylate frameworks. Acc. Chem. Res. 2001, 34, 319–330. [Google Scholar] [CrossRef] [PubMed]
  2. Phan, A.; Doonan, C.J.; Uribe-Romo, F.J.; Knobler, C.B.; O’keeffe, M.; Yaghi, O.M. Synthesis, structure, and carbon dioxide capture properties of zeoliticimidazolate frameworks. Acc. Chem. Res. 2010, 43, 58–67. [Google Scholar] [CrossRef] [PubMed]
  3. Das, M.C.; Xiang, S.; Zhang, Z.; Chen, B. Funktionelle gemischtmetall-organische gerüste mit metalloliganden. Angew. Chem. Int. Ed. 2011, 123, 10696–10707. [Google Scholar] [CrossRef]
  4. Chang, B.K.; Bristowe, N.C.; Bristowe, P.D.; Cheetham, A.K. Van der waals forces in the perfluorinated metal-organic framework zinc 1,2-bis(4-pyridyl)ethane tetrafluoroterephthalate. Phys. Chem. Chem. Phys. 2012, 14, 7059–7064. [Google Scholar] [CrossRef] [PubMed]
  5. Kreno, L.E.; Leong, K.; Farha, O.K.; Allendorf, M.; Van Duyne, R.P.; Hupp, J.T. Metal-organic framework materials as chemical sensors. Chem. Rev. 2012, 112, 1105–1125. [Google Scholar] [CrossRef] [PubMed]
  6. Li, J.R.; Sculley, J.; Zhou, H.C. Metal-organic frameworks for separations. Chem. Rev. 2012, 112, 869–932. [Google Scholar] [CrossRef] [PubMed]
  7. Liu, C.S.; Pilania, G.; Wang, C.; Ramprasad, R. How critical are the van der Waals interactions in polymer crystals? J. Phys. Chem. A 2012, 116, 9347–9352. [Google Scholar] [CrossRef] [PubMed]
  8. Xuan, W.; Zhu, C.; Liu, Y.; Cui, Y. Mesoporous metal-organic framework materials. Chem. Soc. Rev. 2012, 41, 1677–1695. [Google Scholar] [CrossRef] [PubMed]
  9. Furukawa, H.; Cordova, K.E.; O’Keeffe, M.; Yaghi, O.M. The chemistry and applications of metal-organic frameworks. Science 2013, 341, 1230444. [Google Scholar] [CrossRef] [PubMed]
  10. Rosi, N.L.; Eckert, J.; Eddaoudi, M.; Vodak, D.T.; Kim, J.; O’Keeffe, M.; Yaghi, O.M. Hydrogen storage in microporous metal-organic frameworks. Science 2003, 300, 1127–1129. [Google Scholar] [CrossRef] [PubMed]
  11. Belof, J.L.; Stern, A.C.; Eddaoudi, M.; Space, B. On the mechanism of hydrogen storage in a metal-organic framework material. J. Am. Chem. Soc. 2007, 129, 15202–15210. [Google Scholar] [CrossRef] [PubMed]
  12. Bae, Y.S.; Snurr, R.Q. Development and evaluation of porous materials for carbon dioxide separation and capture. Angew. Chem. Int. Ed. 2011, 50, 11586–11596. [Google Scholar] [CrossRef] [PubMed]
  13. Suh, M.P.; Park, H.J.; Prasad, T.K.; Lim, D.W. Hydrogen storage in metal-organic frameworks. Chem. Rev. 2012, 112, 782–835. [Google Scholar] [CrossRef] [PubMed]
  14. Chen, B.; Liang, C.; Yang, J.; Contreras, D.S.; Clancy, Y.L.; Lobkovsky, E.B.; Yaghi, O.M.; Dai, S. A microporous metal-organic framework for gas-chromatographic separation of alkanes. Angew. Chem. Int. Ed. 2006, 45, 1390–1393. [Google Scholar] [CrossRef] [PubMed]
  15. Bae, Y.S.; Mulfort, K.L.; Frost, H.; Ryan, P.; Punnathanam, S.; Broadbelt, L.J.; Hupp, J.T.; Snurr, R.Q. Separation of CO2 from CH4 using mixed-ligand metal-organic frameworks. Langmuir 2008, 24, 8592–8598. [Google Scholar] [CrossRef] [PubMed]
  16. Bloch, E.D.; Queen, W.L.; Krishna, R.; Zadrozny, J.M.; Brown, C.M.; Long, J.R. Hydrocarbon separations in a metal-organic framework with open iron(II) coordination sites. Science 2012, 335, 1606–1610. [Google Scholar] [CrossRef] [PubMed]
  17. Bloch, W.M.; Babarao, R.; Hill, M.R.; Doonan, C.J.; Sumby, C.J. Post-synthetic structural processing in a metal-organic framework material as a mechanism for exceptional CO2/N2 selectivity. J. Am. Chem. Soc. 2013, 135, 10441–10448. [Google Scholar] [CrossRef] [PubMed]
  18. Yoon, M.; Srirambalaji, R.; Kim, K. Homochiral metal-organic frameworks for asymmetric heterogeneous catalysis. Chem. Rev. 2012, 112, 1196–1231. [Google Scholar] [CrossRef] [PubMed]
  19. Genna, D.T.; Wong-Foy, A.G.; Matzger, A.J.; Sanford, M.S. Heterogenization of homogeneous catalysts in metal-organic frameworks via cation exchange. J. Am. Chem. Soc. 2013, 135, 10586–10589. [Google Scholar] [CrossRef] [PubMed]
  20. Nepal, B.; Das, S. Sustained water oxidation by a catalyst cage-isolated in a metal-organic framework. Angew. Chem. Int. Ed. 2013, 52, 7224–7227. [Google Scholar] [CrossRef] [PubMed]
  21. Zhu, Q.L.; Li, J.; Xu, Q. Immobilizing metal nanoparticles to metal–organic frameworks with size and location control for optimizing catalytic performance. J. Am. Chem. Soc. 2013, 135, 10210–10213. [Google Scholar] [CrossRef] [PubMed]
  22. Horcajada, P.; Serre, C.; Maurin, G.; Ramsahye, N.A.; Balas, F.; Vallet-Regí, M.; Sebban, M.; Taulelle, F.; Férey, G. Flexible porous metal-organic frameworks for a controlled drug delivery. J. Am. Chem. Soc. 2008, 130, 6774–6780. [Google Scholar] [CrossRef] [PubMed]
  23. Horcajada, P.; Chalati, T.; Serre, C.; Gillet, B.; Sebrie, C.; Baati, T.; Eubank, J.F.; Heurtaux, D.; Clayette, P.; Kreuz, C.; et al. Porous metal-organic-framework nanoscale carriers as a potential platform for drug delivery and imaging. Nat. Mater. 2010, 9, 172–178. [Google Scholar] [CrossRef] [PubMed]
  24. Della Rocca, J.; Liu, D.; Lin, W. Nanoscale metal-organic frameworks for biomedical imaging and drug delivery. Acc. Chem. Res. 2011, 40, 957–968. [Google Scholar] [CrossRef] [PubMed]
  25. Bushuyev, O.S.; Brown, P.; Maiti, A.; Gee, R.H.; Peterson, G.R.; Weeks, B.L.; Hope-Weeks, L.J. Ionic polymers as a new structural motif for high-energy-density materials. J. Am. Chem. Soc. 2012, 134, 1422–1425. [Google Scholar] [CrossRef] [PubMed]
  26. Tao, G.H.; Parrish, D.A.; Shreeve, J.M. Nitrogen-rich 5-(1-methylhydrazinyl)tetrazole and its copper and silver complexes. Inorg. Chem. 2012, 51, 5305–5312. [Google Scholar] [CrossRef] [PubMed]
  27. Bushuyev, O.S.; Peterson, G.R.; Brown, P.; Maiti, A.; Gee, R.H.; Weeks, B.L.; Hope-Weeks, L.J. Metal-organic frameworks (MOFs) as safer, structurally reinforced energetics. Chem. Eur. J. 2013, 19, 1706–1711. [Google Scholar] [CrossRef] [PubMed]
  28. Li, S.; Wang, Y.; Qi, C.; Zhao, X.; Zhang, J.; Zhang, S.; Pang, S. 3D energetic metal-organic frameworks: Synthesis and properties of high energy materials. Angew. Chem. Int. Ed. 2013, 52, 14031–14035. [Google Scholar] [CrossRef] [PubMed]
  29. Wang, S.H.; Zheng, F.K.; Wu, M.F.; Liu, Z.F.; Chen, J.; Guo, G.C.; Wu, A.Q. Hydrothermal syntheses, crystal structures and physical properties of a new family of energetic coordination polymers with nitrogen-rich ligand N-[2-(1H-tetrazol-5-yl)ethyl]glycine. CrystEngComm 2013, 15, 2616–2623. [Google Scholar] [CrossRef]
  30. Wu, B.D.; Zhang, T.L.; Li, Y.L.; Tong, W.C.; Zhou, Z.N.; Zhang, J.G.; Yang, L. Energetic Compounds Based on 4-Amino-1,2,4-triazole (ATZ) and Picrate (PA):[Zn(H2O)6](PA)2·3H2O and [Zn(ATZ)3](PA)2·2.5H2O]n. Z. Anorg. Allg. Chem. 2013, 639, 2209–2215. [Google Scholar] [CrossRef]
  31. Wu, B.D.; Zhou, Z.N.; Li, F.G.; Yang, L.; Zhang, T.L.; Zhang, J.G. Preparation, crystal structures, thermal decompositions and explosive properties of two new high-nitrogen azide ethylenediamine energetic compounds. New J. Chem. 2013, 37, 646–653. [Google Scholar] [CrossRef]
  32. Gao, W.; Liu, X.; Su, Z.; Zhang, S.; Yang, Q.; Wei, Q.; Chen, S.; Xie, G.; Yang, X.; Gao, S. High-energy-density materials with remarkable thermostability and insensitivity: Syntheses, structures and physicochemical properties of Pb(II) compounds with 3-(tetrazol-5-yl) triazole. J. Mater. Chem. A 2014, 2, 11958–11965. [Google Scholar] [CrossRef]
  33. Liu, X.; Yang, Q.; Su, Z.; Chen, S.; Xie, G.; Wei, Q.; Gao, S. 3D high-energy-density and low sensitivity materials: Synthesis, structure and physicochemical properties of an azide–Cu(II) complex with 3,5-dinitrobenzoic acid. RSC Adv. 2014, 4, 16087–16093. [Google Scholar] [CrossRef]
  34. Tong, W.C.; Yang, L.; Wu, B.D.; Zhang, T.L. Preparation, crystal structure, and thermal decomposition of a novel nitrogen-rich compound Zn3(ATZ)6(N3)6 (ATZ = 4-amino-1,2,4-triazole, N% = 61.7%). Z. Anorg. Allg. Chem. 2014, 640, 980–985. [Google Scholar] [CrossRef]
  35. Wu, B.D.; Bi, Y.G.; Li, F.G.; Yang, L.; Zhou, Z.N.; Zhang, J.G.; Zhang, T.L. A novel stable high-nitrogen energetic compound: Copper(II) 1,2-diaminopropane azide. Z. Anorg. Allg. Chem. 2014, 640, 224–228. [Google Scholar] [CrossRef]
  36. Zhang, S.; Liu, X.; Yang, Q.; Su, Z.; Gao, W.; Wei, Q.; Xie, G.; Chen, S.; Gao, S. A new strategy for storage and transportation of sensitive high-energy materials: Guest-dependent energy and sensitivity of 3D metal-organic-framework-based energetic compounds. Chem. Eur. J. 2014, 20, 7906–7910. [Google Scholar] [CrossRef] [PubMed]
  37. Chen, D.; Huang, S.; Zhang, Q.; Yu, Q.; Zhou, X.; Li, H.; Li, J. Two nitrogen-rich Ni(II) coordination compounds based on 5,5’-azotetrazole: Synthesis, characterization and effect on thermal decomposition for RDX, HMX and AP. RSC Adv. 2015, 5, 32872–32879. [Google Scholar] [CrossRef]
  38. Feng, Y.; Liu, X.; Duan, L.; Yang, Q.; Wei, Q.; Xie, G.; Chen, S.; Yang, X.; Gao, S. In situ synthesized 3D heterometallic metal-organic framework (MOF) as a high-energy-density material shows high heat of detonation, good thermostability and insensitivity. Dalton Trans. 2015, 44, 2333–2339. [Google Scholar] [CrossRef] [PubMed]
  39. Liu, X.; Gao, W.; Sun, P.; Su, Z.; Chen, S.; Wei, Q.; Xie, G.; Gao, S. Environmentally friendly high-energy MOFs: Crystal structures, thermostability, insensitivity and remarkable detonation performances. Green Chem. 2015, 17, 831–836. [Google Scholar] [CrossRef]
  40. Liu, X.; Qu, X.; Zhang, S.; Ke, H.; Yang, Q.; Shi, Q.; Wei, Q.; Xie, G.; Chen, S. High-performance energetic characteristics and magnetic properties of a three-dimensional cobalt(II) metal-organic framework assembled with azido and triazole. Inorg. Chem. 2015, 54, 11520–11525. [Google Scholar] [CrossRef] [PubMed]
  41. McDonald, K.A.; Seth, S.; Matzger, A.J. Coordination polymers with high energy density: An emerging class of explosives. Cryst. Growth Des. 2015, 15, 5963–5972. [Google Scholar] [CrossRef]
  42. Qu, X.; Zhang, S.; Yang, Q.; Su, Z.; Wei, Q.; Xie, G.; Chen, S. Silver(I)-based energetic coordination polymers: Synthesis, structure and energy performance. New J. Chem. 2015, 39, 7849–7857. [Google Scholar] [CrossRef]
  43. Xu, C.X.; Zhang, J.G.; Yin, X.; Jin, X.; Li, T.; Zhang, T.L.; Zhou, Z.N. Cd(II) complexes with different nuclearity and dimensionality based on 3-hydrazino-4-amino-1,2,4-triazole. J. Solid State Chem. 2015, 226, 59–65. [Google Scholar] [CrossRef]
  44. Qu, X.N.; Zhang, S.; Wang, B.Z.; Yang, Q.; Han, J.; Wei, Q.; Xie, G.; Chen, S.P. An Ag(I) energetic metal-organic framework assembled with the energetic combination of furazan and tetrazole: Synthesis, structure and energetic performance. Dalton Trans. 2016, 45, 6968–6973. [Google Scholar] [CrossRef] [PubMed]
  45. Yang, Q.; Song, X.; Ge, J.; Zhao, G.; Zhang, W.; Xie, G.; Chen, S.; Gao, S. A 2D nickel-based energetic MOFs incorporating 3,5-diamino-1,2,4-triazole and malonic acid: Synthesis, crystal structure and thermochemical study. J. Chem. Thermodyn. 2016, 92, 132–138. [Google Scholar] [CrossRef]
  46. Yue, Q.Y.; Lu, Y.M.; Chuan, F.X.; Yuan, D.; Chen, D.Y.; Yang, G.W.; Li, Q.Y. Synthesis, crystal structure, luminescence and thermal behavior of a new energetic zinc(II) compound. Inorg. Chem. Commun. 2016, 68, 68–71. [Google Scholar] [CrossRef]
  47. Zhang, H.; Zhang, M.; Lin, P.; Malgras, V.; Tang, J.; Alshehri, S.M.; Yamauchi, Y.; Du, S.; Zhang, J. A highly energetic N-rich metal-organic framework as a new high-energy-density material. Chem. Eur. J. 2016, 22, 1141–1145. [Google Scholar] [CrossRef] [PubMed]
  48. Zhang, J.; Du, Y.; Dong, K.; Su, H.; Zhang, S.; Li, S.; Pang, S. Taming dinitramide anions within an energetic metal–organic framework: A new strategy for synthesis and tunable properties of high energy materials. Chem. Mater. 2016, 28, 1472–1480. [Google Scholar] [CrossRef]
  49. Zhang, Q.; Shreeve, J.M. Metal-organic frameworks as high explosives: A new concept for energetic materials. Angew. Chem. Int. Ed. 2014, 53, 2540–2542. [Google Scholar] [CrossRef] [PubMed]
  50. Zhang, S.; Liu, X.; Liu, B.; Xia, Z.; Wang, W.; Yang, Q.; Ke, H.; Wei, Q.; Xie, G.; Chen, S. 3D Co(II, III) mixed-valence metal-organic framework affording field-induced slow magnetic relaxation. Sci. China Chem. 2015, 58, 1032–1038. [Google Scholar] [CrossRef]
  51. Zhang, S.; Yang, Q.; Liu, X.; Qu, X.; Wei, Q.; Xie, G.; Chen, S.; Gao, S. High-energy metal–organic frameworks (HE-MOFs): Synthesis, structure and energetic performance. Coord. Chem. Rev. 2016, 307, 292–312. [Google Scholar] [CrossRef]
  52. Wang, Y.; Zhang, J.; Su, H.; Li, S.; Zhang, S.; Pang, S. A simple method for the prediction of the detonation performances of metal-containing explosives. J. Phys. Chem. A 2014, 118, 4575–4581. [Google Scholar] [CrossRef] [PubMed]
  53. Zhang, J.; Shreeve, J.M. 3D nitrogen-rich metal-organic frameworks: Opportunities for safer energetics. Dalton Trans. 2016, 45, 2363–2368. [Google Scholar] [CrossRef] [PubMed]
  54. Deblitz, R.; Hrib, C.G.; Blaurock, S.; Jones, P.G.; Plenikowski, G.; Edelmann, F.T. Explosive werner-type cobalt(III) complexes. Inorg. Chem. Front. 2014, 1, 621–640. [Google Scholar] [CrossRef]
  55. Evers, J.; Gospodinov, I.; Joas, M.; Klapotke, T.M.; Stierstorfer, J. Cocrystallization of photosensitive energetic copper(II) perchlorate complexes with the nitrogen-rich ligand 1,2-di(1H-tetrazol-5-yl)ethane. Inorg. Chem. 2014, 53, 11749–11756. [Google Scholar] [CrossRef] [PubMed]
  56. Blair, L.H.; Colakel, A.; Vrcelj, R.M.; Sinclair, I.; Coles, S.J. Metal-organic fireworks: MOFs as integrated structural scaffolds for pyrotechnic materials. Chem. Commun. 2015, 51, 12185–12188. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Friedrich, M.; Gálvez-Ruiz, J.C.; Klapötke, T.M.; Mayer, P.; Weber, B.; Weigand, J.J. BTA copper complexes. Inorg. Chem. 2005, 44, 8044–8052. [Google Scholar] [CrossRef] [PubMed]
  58. Klapötke, T.M.; Mayer, P.; Stierstorfer, J.; Weigand, J.J. Bistetrazolylamines—Synthesis and characterization. J. Mater. Chem. 2008, 18, 5248. [Google Scholar] [CrossRef]
  59. Guo, Y.; Gao, H.; Twamley, B.; Shreeve, J.M. Energetic nitrogen rich salts of N,N-bis[1(2)H-tetrazol-5-yl]amine. Adv. Mater. 2007, 19, 2884–2888. [Google Scholar] [CrossRef]
  60. Guo, Y.; Tao, G.H.; Zeng, Z.; Gao, H.; Parrish, D.A.; Shreeve, J.M. Energetic salts based on monoanions of N,N-bis[1(2)H-tetrazol-5-yl]amine. Chem. Eur. J. 2010, 16, 3753–3762. [Google Scholar] [CrossRef] [PubMed]
  61. Dong, L.L.; He, L.; Liu, H.Y.; Tao, G.H.; Nie, F.D.; Huang, M.; Hu, C.W. Nitrogen-rich energetic ionic liquids based on the N,N-bis[1H-tetrazol-5-yl]amine anion—Syntheses, structures, and properties. Eur. J. Inorg. Chem. 2013, 5009–5019. [Google Scholar] [CrossRef]
  62. Xue, B.D.; Yang, Q.; Chen, S.P.; Gao, S.L. Synthesis, crystal structure, and thermodynamics of a high-nitrogen copper complex with N,N-bis[1(2)H-tetrazol-5-yl] amine. J. Therm. Anal. Calorim. 2010, 101, 997–1002. [Google Scholar] [CrossRef]
  63. Li, N.; Chen, W.B.; Guan, Y.F.; OuYang, Z.J.; Dong, W. Chlorine anion–π and ππ interactions in two tetrazolyl derivative based Cu2+ complexes and quantum chemical calculations. Inorg. Chim. Acta 2014, 409, 349–352. [Google Scholar] [CrossRef]
  64. Qu, B.; Lu, X.; Wu, Y.; You, X.; Xu, X. Synthesis of copper micro-rods with layered nano-structure by thermal decomposition of the coordination complex Cu(BTA)2. Nanoscale Res. Lett. 2015, 10, 1–5. [Google Scholar] [CrossRef] [PubMed]
  65. Cook, C.; Habib, F.; Aharen, T.; Clerac, R.; Hu, A.; Murugesu, M. High-temperature spin crossover behavior in a nitrogen-rich Fe(III)-based system. Inorg. Chem. 2013, 52, 1825–1831. [Google Scholar] [CrossRef] [PubMed]
  66. Liu, N.; Yue, Q.; Wang, Y.Q.; Cheng, A.L.; Gao, E.Q. Coordination compounds of bis(5-tetrazolyl)amine with manganese(II), zinc(II) and cadmium(II): Synthesis, structure and magnetic properties. Dalton Trans. 2008, 34, 4621–4629. [Google Scholar] [CrossRef] [PubMed]
  67. Lu, Y.B.; Wang, M.S.; Zhou, W.W.; Xu, G.; Guo, G.C.; Huang, J.S. Novel 3-D PtS-like tetrazolate-bridged manganese(II) complex exhibiting spin-canted antiferromagnetism and field-induced spin-flop transition. Inorg. Chem. 2008, 47, 8935–8942. [Google Scholar] [CrossRef] [PubMed]
  68. Jiang, T.; Zhang, X.M. Temperature tuned synthesis of zinc N,N-bis[1(2)H-tetrazol-5-yl]-amine complexes: From Zn3O cluster-based 3-connected 63-hcb and (3,6)-connected (43)2(46)-kgd layers to Zn5(OH)4 chain-based 3D open framework. Cryst. Growth Des. 2008, 8, 3077–3083. [Google Scholar] [CrossRef]
  69. Li, F.; Bi, Y.; Zhao, W.; Zhang, T.; Zhou, Z.; Yang, L. Nitrogen-rich salts based on the energetic [monoaquabis N,N-bis[1(2)H-tetrazol-5-yl] amine)-zinc(II)] anion: A promising design in the development of new energetic materials. Inorg. Chem. 2015, 54, 2050–2057. [Google Scholar] [CrossRef] [PubMed]
  70. Lin, J.M.; Guan, Y.F.; Wang, D.Y.; Dong, W.; Wang, X.T.; Gao, S. Syntheses, structures and properties of seven isomorphous 1D Ln3+ complexes Ln(BTA)(HCOO)(H2O)3 (H2BTA = bis(tetrazoly)amine, Ln = Pr, Gd, Eu, Tb, Dy, Er, Yb) and two 3D Ln3+ complexes Ln(HCOO)3 (Ln = Pr, Nd). Dalton Trans. 2008, 44, 6165–6169. [Google Scholar] [CrossRef] [PubMed]
  71. Wang, W.; Chen, S.; Gao, S. Syntheses and characterization of lead(II) N,N-bis[1(2)H-tetrazol-5-yl] amine compounds and effects on thermal decomposition of ammonium perchlorate. Eur. J. Inorg. Chem. 2009, 2009, 3475–3480. [Google Scholar] [CrossRef]
  72. Wu, J.T.; Zhang, J.G.; Yin, X.; He, P.; Zhang, T.L. Synthesis and characterization of the nitrophenol energetic ionic salts of 5,6,7,8-tetrahydrotetrazolo[1,5-b][1,2,4]triazine. Eur. J. Inorg. Chem. 2014, 2014, 4690–4695. [Google Scholar] [CrossRef]
  73. Liu, Q.; Jin, B.; Peng, R.; Guo, Z.; Zhao, J.; Zhang, Q.; Shang, Y. Synthesis, characterization and properties of nitrogen-rich compounds based on cyanuric acid: A promising design in the development of new energetic materials. J. Mater. Chem. A 2016, 4, 4971–4981. [Google Scholar] [CrossRef]
  74. Allen, F.H.; Kennard, O.; Watson, D.G.; Brammer, L.; Orpen, A.G.; Taylor, R. Tables of bond lengths determined by X-ray and neutron diffraction. Part I. Bond lengths in organic compounds. J. Chem. Soc. Perkin Trans. 2 1987, S1–S19. [Google Scholar] [CrossRef]
  75. Li, F.; Zhao, W.; Chen, S.; Zhang, T.; Zhou, Z.; Yang, L. Nitrogen-rich alkali metal salts (Na and K) of [bis(N,N-bis[1H-tetrazol-5-yl] amine)-zinc(II)] anion: Syntheses, crystal structures, and energetic properties. Z. Anorg. Allg. Chem. 2015, 641, 911–916. [Google Scholar] [CrossRef]
  76. Cacho-Bailo, F.; Téllez, C.; Coronas, J. Interactive Thermal Effects on Metal-Organic Framework Polymer Composite Membranes. Chem. Eur. J. 2016, 22, 9533–9536. [Google Scholar] [CrossRef] [PubMed]
  77. Kissinger, H.E. Reaction kinetics in differential thermal analysis. Anal. Chem. 1957, 29, 1702–1706. [Google Scholar] [CrossRef]
  78. Doyle, C.D. Kinetic analysis of thermogravimetric data. J. Appl. Polym. Sci. 1961, 5, 285–292. [Google Scholar] [CrossRef]
  79. Ozawa, T. A new method of analyzing thermogravimetric data. Bull. Chem. Soc. Jpn. 1965, 38, 1881–1886. [Google Scholar] [CrossRef]
  80. Geetha, M.; Nair, U.R.; Sarwade, D.B.; Gore, G.M.; Asthana, S.N.; Singh, H.J. Studies on CL-20: The most powerful high energy material. J. Therm. Anal. Calorim. 2003, 73, 913–922. [Google Scholar] [CrossRef]
  81. Turcotte, R.; Vachon, M.; Kwok, Q.S.; Wang, R.; Jones, D.E. Thermal study of HNIW (CL-20). Thermochim. Acta 2005, 433, 105–115. [Google Scholar] [CrossRef]
  82. Sheldrick, G.M. SHELXS-97, Program for Crystal Structure Solution; University of Göttingen: Göttingen, Germany, 1997. [Google Scholar]
  83. Sheldrick, G.M. SHELXS-97, Program for Crystal Structure Refinement; University of Göttingen: Göttingen, Germany, 1997. [Google Scholar]
  84. Delley, B. An all-electron numerical method for solving the local density functional for polyatomicmolecules. J. Chem. Phys. 1990, 92, 508–517. [Google Scholar] [CrossRef]
  85. Monkhorst, H.J.; Pack, J.D. Special points for Brillouin-zone integrations. Phys. Rev. B 1976, 13, 5188–5192. [Google Scholar] [CrossRef]
  86. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Design strategy for 1D, 2D, and 3D energetic MOFs (Reprinted from Refs. [28,49]. (Copyright 2013 and 2014 Wiley).
Figure 1. Design strategy for 1D, 2D, and 3D energetic MOFs (Reprinted from Refs. [28,49]. (Copyright 2013 and 2014 Wiley).
Materials 09 00681 g001
Figure 2. 1D energetic MOFs of nickel hydrazine-perchlorate (NHP, left) and nickel hydrazine-nitrate (NHN, right) (Reprinted from Ref. [25]. Copyright 2012 American Chemical Society).
Figure 2. 1D energetic MOFs of nickel hydrazine-perchlorate (NHP, left) and nickel hydrazine-nitrate (NHN, right) (Reprinted from Ref. [25]. Copyright 2012 American Chemical Society).
Materials 09 00681 g002
Scheme 1. Synthesis of energetic 3D MOFs [Pb(bta)·2H2O]n.
Scheme 1. Synthesis of energetic 3D MOFs [Pb(bta)·2H2O]n.
Materials 09 00681 sch001
Figure 3. Ball-and-stick molecular structure of [Pb(bta)·2H2O] (a) and coordination mode of ligand (b).
Figure 3. Ball-and-stick molecular structure of [Pb(bta)·2H2O] (a) and coordination mode of ligand (b).
Materials 09 00681 g003
Figure 4. Ball-and-stick packing diagram of [Pb(bta)·2H2O]n viewed down the a-axis(a); b-axis (b); and a–c diagonal (c).
Figure 4. Ball-and-stick packing diagram of [Pb(bta)·2H2O]n viewed down the a-axis(a); b-axis (b); and a–c diagonal (c).
Materials 09 00681 g004
Figure 5. Coordination polyhedron geometry of [Pb(bta)·2H2O]n (a) polyhedrons with Pb(II) as the center viewed down a-axis (b) and the a–c diagonal (c).
Figure 5. Coordination polyhedron geometry of [Pb(bta)·2H2O]n (a) polyhedrons with Pb(II) as the center viewed down a-axis (b) and the a–c diagonal (c).
Materials 09 00681 g005
Figure 6. DSC curve of [Pb(bta)·2H2O]n at a heating rate of 5 K·min−1.
Figure 6. DSC curve of [Pb(bta)·2H2O]n at a heating rate of 5 K·min−1.
Materials 09 00681 g006
Figure 7. TGA curve of [Pb(bta)·2H2O]n at a heating rate of 5 K·min−1.
Figure 7. TGA curve of [Pb(bta)·2H2O]n at a heating rate of 5 K·min−1.
Materials 09 00681 g007
Figure 8. DSC curves of [Pb(bta)·2H2O]n at different heating rate.
Figure 8. DSC curves of [Pb(bta)·2H2O]n at different heating rate.
Materials 09 00681 g008
Figure 9. Bar chart representation of ∆Hdet values in literature for common explosive materials and previously reported values for energetic MOFs along with the predicted ∆Hdet value for [Pb(bta)·2H2O]n. The errorbars correspond to the 96% statistical-confidence level for these values.
Figure 9. Bar chart representation of ∆Hdet values in literature for common explosive materials and previously reported values for energetic MOFs along with the predicted ∆Hdet value for [Pb(bta)·2H2O]n. The errorbars correspond to the 96% statistical-confidence level for these values.
Materials 09 00681 g009
Table 1. Crystal data and structure refinement data of [Pb(bta)·2H2O]n.
Table 1. Crystal data and structure refinement data of [Pb(bta)·2H2O]n.
Empirical FormulaC2H5N9O2Pb
Formula weight394.34
Crystal Colorcolorless
Crystal size (mm3)0.21 × 0.20 × 0.19 mm
Crystal systemMonoclinic
Space groupP21/n
a (Å)6.592(5)
b (Å)11.987(9)
c (Å)10.552(8)
α (°)90
β (°)104.856(12)
γ (°)90
V3)806.0(10)
Z4
ρcalcd (g·cm−3)3.250
T (K)150(2)
F (000)712
θ (°)2.62 to 25.00
R int.0.0425
Data1423
Restraints4
parameters127
GOF a on F21.051
R1 b (I > 2σ (I))0.0241
ωR2 c (I > 2σ (I))0.0655
R1 (all data)0.0261
a GOF = Goodness of Fit; b R1 = ∑||Fo| − |Fc||/∑|Fo|; c ωR2 = [(ω(Fo2Fc2)2)/ω(Fo2)2]1/2.
Table 2. The calculated kinetic parameters for the exothermic decomposition processes of [Pb(bta)·2H2O]n.
Table 2. The calculated kinetic parameters for the exothermic decomposition processes of [Pb(bta)·2H2O]n.
β/(K·min−1)Tp/(K)KissingerOzawa-Doyle
E/(kJ·mol−1)ln ArE/(kJ·mol−1)r
5614.9436.184.930.9984424.60.9984
10620.1
15623.0
20624.7
Table 3. Physicochemical properties of [Pb(bta)·2H2O]n.
Table 3. Physicochemical properties of [Pb(bta)·2H2O]n.
Td ap bN cΔHdet dD eP fIS g
[Pb(bta)·2H2O]n3423.25031.984.978.96347.47>40.0
CHP h [25]1941.94814.715.238.22531.730.5
CHHP h [27]2312.00028.253.146.20517.960.8
ZnHHP h [27]2932.11723.612.937.01623.582.5
ATRZ-1 h [29]2431.68053.3515.149.16035.6822.5
ATRZ-2 h [29]2572.16043.765.787.77329.7030.0
[Pb(Htztr)2 h (H2O)]n [32]3402.51939.405.697.71531.57>40.0
[Pb(Htztr)(O)]n h [32]3183.51127.200.948.12240.12>40.0
HMX [25]2871.95037.845.528.90038.397.4
RDX [25]2101.80637.805.808.60033.927.4
a Decomposition temperature; b Density from X-ray diffraction analysis (g·cm−3); c Nitrogen content (%); d The heat of detonation (kJ·g−1); e Detonation velocity (km·s−1); f Detonation pressure (GPa); g Impact sensitivity; h CHP = cobalt hydrazine perchlorate; CHHP = cobalt hydrazine hydrazinecarboxylate perchlorate; ZnHHP = zinc hydrazine hydrazinecarboxylate perchlorate; ATRZ = 4,4’-azo-1,2,4-triazole; Htztr = 3-(tetrazol-5-yl)triazole.

Share and Cite

MDPI and ACS Style

Liu, Q.; Jin, B.; Zhang, Q.; Shang, Y.; Guo, Z.; Tan, B.; Peng, R. Nitrogen-Rich Energetic Metal-Organic Framework: Synthesis, Structure, Properties, and Thermal Behaviors of Pb(II) Complex Based on N,N-Bis(1H-tetrazole-5-yl)-Amine. Materials 2016, 9, 681. https://doi.org/10.3390/ma9080681

AMA Style

Liu Q, Jin B, Zhang Q, Shang Y, Guo Z, Tan B, Peng R. Nitrogen-Rich Energetic Metal-Organic Framework: Synthesis, Structure, Properties, and Thermal Behaviors of Pb(II) Complex Based on N,N-Bis(1H-tetrazole-5-yl)-Amine. Materials. 2016; 9(8):681. https://doi.org/10.3390/ma9080681

Chicago/Turabian Style

Liu, Qiangqiang, Bo Jin, Qingchun Zhang, Yu Shang, Zhicheng Guo, Bisheng Tan, and Rufang Peng. 2016. "Nitrogen-Rich Energetic Metal-Organic Framework: Synthesis, Structure, Properties, and Thermal Behaviors of Pb(II) Complex Based on N,N-Bis(1H-tetrazole-5-yl)-Amine" Materials 9, no. 8: 681. https://doi.org/10.3390/ma9080681

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop