Yilin
Jiang
,
Jinlin
Yin
,
Ruonan
Xi
and
Honghan
Fei
*
Shanghai Key Laboratory of Chemical Assessment and Sustainability, School of Chemical Science and Engineering, Tongji University, 1239 Siping Rd., Shanghai 200092, China. E-mail: fei@tongji.edu.cn
First published on 19th January 2024
Crystal engineering of metal halide hybrids is critical to investigate their structure–property relationship and advance their photophysical applications, but there have been limited efforts to employ coordination chemistry to precisely control the dimensionality of metal halide sublattices. Herein, we present a coordination–assembly synthetic strategy developed for the rational modulation of lead halide dimensionality, realizing the transition from 2D to 3D architectures. This manipulation is achieved by utilizing three organocarboxylates featuring the identical cyclohexane backbone unit. Specifically, the 1,4-cyclohexanedicarboxylate and 1,2,4,5-cyclohexanetetracarboxylate ligands facilitate the formation of quasi-2D layered structures, characterized by weakly corrugated and strongly corrugated lead halide layers, respectively. Importantly, the introduction of the 1,2,3,4,5,6-cyclohexanehexacarboxylate ligand results in coordination architectures featuring 3D lead chloride/bromide sublattices. The formation of the 3D coordination architectures templated by the 1,2,3,4,5,6-cyclohexanehexacarboxylate ligand affords extended wavelength coverage and superior carrier transport properties compared to their quasi-2D layered analogues. Importantly, both the 2D and 3D lead halide-based coordination polymers exhibit high aqueous stability over a wide pH range, outperforming the conventional ionic-bound lead halides. Notably, the chemically stable 3D lead bromide exhibits efficient photocatalytic ethylbenzene oxidation with the conversion rate of 498 μmol g−1 h−1, substantially higher than its 2D lead bromide counterparts. This work highlights the important role of coordination chemistry in the rational design of metal halide hybrids, which is crucial for advancing their photophysical properties and applications.
A variety of synthetic strategies have been developed to control the dimensionality of lead halide sublattices at the molecular level.16 For example, incorporating organoammonium cations with large molecular size and high rigidity often leads to the formation of low-dimensional lead halide hybrids.17–19 In addition, a commonly employed approach involves varying synthesis conditions such as precursor ratios, temperatures and solvents, which often serendipitously result in the formation of lead halide hybrids with different dimensionalities.20–24 More recently, the inclusion of guest molecules (e.g. H2O, NH3 and MeOH) into the parent lattice of lead halides has been discovered as a successful approach to isolate the inorganic halide connectivity, leading to the formation of low-dimensional lead halide structures.25–28 Although less explored in the literature, rational synthetic approaches based on coordination chemistry to modulate the dimensionality of lead halides remain limited.29
Our group focused on the synthesis of lead halide-based coordination polymers, which are extended frameworks consisting of lead halide inorganic components and coordinating anionic ligands as the organic components.30–33 By employing this general synthetic strategy, we have significantly advanced the structural integrity and intrinsic stability of lead halide hybrids, primarily attributed to the formation of 3D coordination networks instead of ionically bound structures.34–36 This class of extended coordinative architectures display excellent aqueous stability across a wide pH range and even under boiling conditions.30,37 However, the vast majority of these structures adopt quasi-2D layered arrangements with tunable layer corrugation and thickness.30,38,39 It is challenging to achieve the rational transition from 2D to 3D in lead halide sublattices for this class of coordination polymers, but it is a compelling pursuit to discover the 3D analogues combining the outstanding photophysical properties and high aqueous stability.
Herein, we have successfully achieved precise control over the dimensionality, transitioning from 2D to 3D lead halide sublattices for the coordination polymers. This was realized by employing three organocarboxylates that share the identical cyclohexane backbone unit. The use of 1,4-cyclohexanedicarboxylate (cdc2−) and 1,2,4,5-cyclohexanetetracarboxylate (ctc4−) led to the formation of quasi-2D layered structures with weakly corrugated and strongly corrugated features, respectively. Intriguingly, the incorporation of 1,2,3,4,5,6-cyclohexanehexacarboxylate (chc6−) gives rise to two chemically stable 3D lead chloride/bromide coordination architectures that exhibit remarkable aqueous pH stability. Moreover, our investigation involving band structure studies, Hall effect measurements and ultrafast transient absorption (TA) spectroscopy indicates that the 3D structures templated by chc ligands demonstrate wide wavelength coverage and enhanced carrier-transport characteristics compared to their quasi-2D analogues. In addition, since developing C(sp3)–H activation under mild conditions is crucial for direct conversion of abundant hydrocarbons,8,9,40,41 we have employed the photocatalytic benzyl C(sp3)–H oxidation as a model reaction to evaluate their photocatalytic performance. Notably, the 3D lead bromide coordination polymer loaded with trace amount of the Pt co-catalyst exhibits the highest ethylbenzene conversion rates of 498 μmol g−1 h−1 with a selectivity of 61%, substantially outperforming the 2D lead bromide counterparts.
Despite the pronounced corrugation observed in TJU-13, the interlayer spacing of ∼13 Å was unfavorable for efficient carrier transport. Therefore, the lead halide dimensionality modulation from 2D to 3D is further investigated by using chc6− as the templating ligand. Solvothermal reaction of PbCl2 and chc in deionized H2O afforded colorless block-shaped crystals of [Pb12Cl17(OH)]6+(chc6−), denoted as TJU-14(Cl). X-ray crystallography studies revealed that TJU-14(Cl) is crystallized in the trigonal crystal system with the space group of Pc1 (Table S2†). TJU-14(Cl) comprises edge- and vertex-sharing PbCl5 and PbCl6 units extending in three dimensions, defining a 3D inorganic [Pb12Cl17(OH)]6+ framework (Fig. 1e and S3†). The 3D inorganic hybrid features an array of honeycomb-like channels along the c-axis, with an open aperture diameter of ∼9.4 Å but accommodating the chc6− ligands. Notably, all six carboxylates in the chc6− ligands coordinatively bonded with the Pb2+ centers, forming an array of 12-membered-ring channels. This phenomenon suggests the crucial role of an increasing number of carboxylate groups in organic ligands for the 2D to 3D dimensionality modulation. Both crystallographically independent Pb2+ centers adopt eight-coordinate geometries, with one coordinating to five Cl− anions, two carboxylate oxygens and one μ3-OH, while the other coordinating with six Cl− anions and two carboxylate oxygens.
The substitution of PbCl2 with PbBr2 in our synthetic protocol resulted in the formation of a coordination polymer with a more densely packed 3D lead bromide topology ([Pb15Br24]6+(chc6−), namely TJU-15(Br)). TJU-15(Br) crystallizes in the highly symmetric cubic crystal system with the Pa-3 space group (Table S2†). TJU-15(Br) also consists of edge- and vertex-sharing PbBr5 and PbBr6 units extending in three dimensions to define the inorganic framework and the chc ligand coordinatively bridge the lead centers within the pore cages (Fig. 1f). The inorganic sublattice of TJU-15(Br) exhibits higher overall occupancy and a more densely packed structure compared to TJU-14(Cl). Each chc6− ligand comprises 12 carboxylate oxygens, all of which form coordination bonds with the Pb2+ centers. This arrangement results in the effective spatial separation of individual ligands within the inorganic skeleton. Half of the crystallographically independent Pb atoms exhibit coordination to six bridging Br atoms, forming the [PbBr6]4− octahedron as for perovskites, while the other Pb atoms adopt a seven-coordinate environment with two carboxylate oxygens and five bromides. The Pb–Br bond distances (2.920–3.310 Å) are well within the accepted range for a covalent Pb–Br bond, suggesting that TJU-15(Br) indeed occupies a 3D lead bromide sublattice.
All of the four new lead halide coordination polymers in this work, TJU-13(Cl), TJU-13(Br), TJU-14(Cl), and TJU-15(Br), can be synthesized with high yields (>75%) and high phase purity, evidenced by C/H elemental analysis and powder X-ray diffraction (PXRD) patterns. All of the experimental PXRD matched well with the theoretical patterns simulated from single-crystal data (Fig. S4†). Thermogravimetric analysis (TGA) and ex situ thermodiffraction suggest the thermal stability of TJU-13(Cl) and TJU-13(Br) up to 120 °C (Fig. S5 and S6†). The increased lead halide dimensionality and higher charge of chc6− ligands synergistically render the improved thermal stability up to 200 °C for TJU-14(Cl) and 220 °C for TJU-15(Br), respectively (Fig. 2a and b and S5†). The chemical stability studies were performed by incubating the as-synthesized crystals in aqueous solutions over a wide range of pH (4–10) at room temperature for 24 h. All of our four coordination polymers retain high crystallinity and demonstrate negligible loss in mass balance, superior to the conventional ionically bound lead halide hybrids (Fig. 2a, b and S6†).
The band structures of TJU-14(Cl) and TJU-15(Br) were further studied by the density functional theory (DFT) calculations using the Perdew–Burke–Ernzerhof (PBE) functional, which demonstrated the calculated bandgaps to be 3.45 eV for TJU-14(Cl) and 2.97 eV for TJU-15(Br), respectively (Fig. 2e and f). Both values matched well with the experimental bandgaps (Fig. 2c and d). In addition, TJU-14(Cl) possessed an indirect bandgap configuration, with the conductive band minimum (CBM) located at G point and the valence band maximum (VBM) at M point, agreeing with the experimental Tauc plot (Fig. 2e and S7†). Meanwhile, both frontier orbitals of TJU-15(Br) were located at G point, suggesting the direct bandgap nature (Fig. 2f and S8†). Furthermore, the calculated projected density of states (pDOS) of both coordination polymers showed that the VBMs were mainly contributed by the Cl 3p orbitals for TJU-14(Cl) or the Br 4p orbitals for TJU-15(Br), which elucidated the broader absorption wavelength coverage for the bromide over its chloride counterpart (Fig. 2e and f). Meanwhile, the ligands partially contribute to the VBM, owing to the Pb–carboxylate coordinative behavior. The CBMs of both materials were primarily dominated by the Pb 6p orbitals, analogous to many ionic bound lead halide hybrids.
Ultraviolet photoelectron spectroscopy (UPS) suggested the ionization energies, defined as the energy differences between the VBMs and the vacuum levels, to be 6.11 eV for TJU-13(Br) and 6.64 eV for TJU-15(Br) (Fig. S9†). The energy level positions of their VBMs were determined to be 1.96 eV for TJU-13(Br) and 2.55 eV for TJU-15(Br), respectively, both of which were measured below the Fermi level (Ef). After converting the energy potentials into volts with respect to the normal hydrogen electrode (NHE), the frontier orbitals of TJU-13(Br) were found to be 1.61 V for VBM and −1.82 V for CBM (vs. NHE, pH = 7) while the VBM and CBM of TJU-15(Br) are 2.14 V and −1.15 V (vs. NHE, pH = 7) (Fig. S10†). The CBMs of both materials were more negative than the redox potentials of O2/˙O2− (−0.28 V vs. NHE, pH = 7), indicating their potentials for aerobic photooxidation of ethylbenzene.
AC Hall effect measurement at room temperature provided further insights into the carrier dynamics of the materials. We have measured the carrier mobility of three independent samples at different current densities from 10 nA to 1 mA, and reported the mean and error to guarantee the good statistical accuracy. The quasi-2D TJU-13(Br) exhibited a carrier mobility of ∼3.37 ± 0.13 cm2 V−1 s−1 while the 3D bromide counterpart (TJU-15(Br)) demonstrated an enhanced carrier mobility of ∼3.83 ± 0.22 cm2 V−1 s−1, despite both materials having analogous carrier concentrations (Table 1). These values again confirmed that the high-dimensional lead bromide hybrid favors efficient charge separation and transport. The carrier transport length (LD) was calculated to be in the range of 0.45–1.21 μm for TJU-15(Br) based on the equation LD = (kBT/e × μ × τ)1/2, where kB is the Boltzmann's constant, e is the electron charge, T is the absolute temperature, μ is the Hall carrier mobility, and τ is the carrier lifetime.
Material | Band gap [eV] | Mobility [cm2 V−1 s−1] | Carrier concentration [cm−3] | TA lifetimes [ps] | PL lifetimes [ns] |
---|---|---|---|---|---|
TJU-13(Br) | 3.43 | 3.37 | 3.96 × 1016 | τ 1 = 108.51 | τ 1 = 9.19 |
τ 2 = 555.77 | τ 2 = 131.35 | ||||
TJU-15(Br) | 3.29 | 3.83 | 2.87 × 1016 | τ 1 = 12.37 | τ 1 = 20.31 |
τ 2 = 334.27 | τ 2 = 148.73 |
Ultrafast femtosecond TA spectroscopy (the pump laser is set as 350 nm, above the band gap, the range of probe wavelength is 400–700 nm) was further employed to gain deeper insight of the carrier transport of TJU-15(Br). The pseudo-color TA spectra of TJU-15(Br) revealed a broad positive photoinduced absorption (PIA) band ranging from 460 to 650 nm, in good agreement with the formation of self-trapped states (Fig. 3c).49,50 The gradually decreased intensity of PIA signals at longer probe delay times (∼1 ns) suggested a detrapping process. The TA dynamics of TJU-15(Br) were fitted using a global bi-exponential equation based on the decay curve, affording two time constants of 12.4 ps and 334.3 ps, respectively (Fig. 3d). The ultrafast time component τ1 represents the rapid trapping of free excitons,51 while the component τ2 is attributed to the nonradiative recombination of the STEs.52,53 Both components were consistent with the STE mechanism, providing further evidence for TJU-15(Br)'s intrinsic large-Stokes shifted broadband emission (Fig. 3a and b). In addition, electrochemical impedance spectroscopy (EIS) and transient photocurrent responses on TJU-13(Br) and TJU-15(Br) were further investigated to compare the charge separation/transfer properties between 2D and 3D structures. The 3D analogue, TJU-15(Br), exhibited a smaller Nyquist plot diameter and higher photocurrent intensity over multiple on–off cycles, strongly demonstrating that the charge transport increased via the modulation of the lead halide dimensionality (Fig. 3e and f).
In order to enhance photogenerated carrier separation and transport, we have performed the photodeposition of Pt nanoparticles on three bromide photocatalysts using K2PtCl4 as the precursor. Inductively coupled plasma-optical emission spectroscopy (ICP-OES) indicated the Pt loadings of 0.43 wt% for TJU-11(Br), 0.35 wt% for TJU-13(Br), and 0.44 wt% for TJU-15(Br), respectively (Table S4†). Importantly, the crystallinity of all three coordination polymers was well retained after the Pt nanoparticle deposition, as evidenced by PXRD (Fig. S14, S15 and S16†). For TJU-15(Br), transmission electron microscopy (TEM) revealed successful photodeposition of Pt nanoparticles with a particle size distribution of 4–12 nm (Fig. S17†). High-resolution TEM images indicated a d-spacing of ∼0.226 nm, corresponding to the (111) lattice planes of face-centered cubic (fcc) Pt (Fig. S17†). Energy-dispersive X-ray spectroscopy (EDS) elemental mapping further confirmed the presence of Pb, Br and Pt in a single crystal of TJU-15(Br) (Fig. S18†).
Among the three lead bromide photocatalysts, Pt0.44@TJU-15(Br) exhibited the highest photocatalytic performances with the ethylbenzene conversion rate of 326 μmol g−1 h−1, similarly superior to the 2D lead bromide analogues (i.e., Pt0.43@TJU-11(Br) and Pt0.35@TJU-13(Br)) (Fig. 4a and Table S3†). In order to investigate the effect of particle size on the catalytic activity, we first tested the particle sizes of the three samples. Fig. S19† shows similar distributional center positions and trends of the three catalysts. This excludes the effect of differences in particle size on catalytic activity.
Importantly, control experiments with the absence of light or photocatalyst resulted in negligible turnover. In order to further verify whether Pt deposition initiates the aerobic oxidation of produced alcohols in the reaction, we have performed the control experiments using 1-phenylethanol as the substrate under the same conditions but in the absence of light. As a result, no acetophenone products were detected, ruling out the possibility of aerobic oxidation initiated by Pt species. Therefore, the modified Pt nanoparticles play a role in promoting rapid charge transfer for photocatalysis (Fig. S20†).54,55 Using PbBr2 as the photocatalyst only yielded a low ethylbenzene conversion rate of 121 μmol g−1 h−1.
The different amounts of Pt loadings on TJU-15(Br) were examined in the ethylbenzene oxidation, and the optimized loading amount was determined to be 0.66 wt% by ICP-OES. Pt0.66@TJU-15(Br) exhibited the highest ethylbenzene conversion rates of 498 μmol g−1 h−1, with a selectivity of 61% for the corresponding ketone (i.e., acetophenone) and a selectivity of 39% for the alcohol (i.e., 1-phenylethanol) (Fig. 4b and c). Long-term stability of the Pt0.66@TJU-15(Br) catalyst was investigated by recovering the used photocatalysts by centrifugation and directly performing in the consecutive catalytic runs. No obvious loss of photocatalytic activity (<8%) was observed for at least five photocatalytic cycles (Fig. 4d). The high recyclability of Pt0.66@TJU-15(Br) was further confirmed by the retention of the crystallinity and negligible leaching of Pb2+ species (∼2%) (Fig. S21†). Moreover, our lead bromide coordination polymers do not require the addition of anhydrous Na2SO4 to remove the water generated during the reaction, distinguishing them from the moisture-sensitive perovskite-type metal halide photocatalysts reported previously.56
The photocatalytic mechanism was investigated by performing control experiments using different radical scavengers (Fig. 4e). The addition of tetra-methylpiperidine N-oxide (TEMPO) as the quenching agent for all radicals completely shut down the turnover, demonstrating that the reaction was primarily governed by a free radical-based process. Using tert-butanol (TBA) as a hydroxyl radical (˙OH) scavenger had minimal effect on the photocatalytic activity, suggesting that ˙OH radicals do not play a significant role in the photocatalysis. Moreover, using 1,4-benzoquinone (BQ) as the superoxide radical (˙O2−) scavenger or K2S2O8 as the electron (e−) scavenger led to a decrease in photocatalytic performances, suggesting that the ˙O2− and electrons are involved in the photocatalytic mechanism. Indeed, the existence of ˙O2− was again confirmed by electron paramagnetic resonance (EPR) spectroscopy under visible-light irradiation, where the characteristic peaks of the DMPO-˙O2− adduct was clearly observed in the photocatalytic reaction solution (Fig. 4f). Both the carbon-centered radical (˙R) scavenger (e.g. butylated hydroxytoluene, BHT) and the hole (h+) scavenger (e.g., (NH4)2C2O4) also render an obvious loss in photocatalytic turnovers, indicating that the carbon-centered radical and the photogenerated holes are also intermediates for the oxidation of ethylbenzene. These radical studies collectively suggest that the activation of C(sp3)–H bonds to generate ˙R radicals is a key reaction step in the mechanistic pathway, followed by the reaction between ˙R and ˙O2−/O2 to yield oxidized products (Fig. S22†). Based on the above experiments, the photocatalysis mechanism is concluded in Fig. S22.† First, photoinduced electrons and holes were generated in the CB and VB of lead bromide photocatalysts under visible light irradiation. Following this, O2 molecules captured the photoelectrons to generate active ˙O2−, whereas the holes drove the challenging oxidation of C(sp3)–H bonds in ethylbenzene to form ˙R radicals. Subsequently, the as-formed ˙R intermediates reacted with ˙O2− to form oxidized products (acetophenone as the major product). The ˙R intermediates may also react with O2 to form oxidized products (acetophenone and 1-phenylethanol are generated in equal proportions).12,57,58
Footnote |
† Electronic supplementary information (ESI) available: Experimental details and additional characterization. CCDC 2294271, 2294272, 2294273 and 2294304. For ESI and crystallographic data in CIF or other electronic format see DOI: https://doi.org/10.1039/d3sc04969h |
This journal is © The Royal Society of Chemistry 2024 |