Ying
Sun
a,
Qian
Ma
b,
Dongheng
Zhao
b,
Pan
Gao
a,
Qi
Wang
a,
Zeyu
Guo
a and
Xiaomei
Jiang
*a
aSchool of Preventive Medicine Sciences (Institute of Radiation Medicine), Shandong First Medical University & Shandong Academy of Medical Sciences, No. 6699 Qingdao Road, Jinan 250117, People's Republic of China. E-mail: jiangxiaomei@sdfmu.edu.cn
bSchool of Material Science and Engineering, University of Jinan, No. 336 Nanxinzhuang West Road, Jinan 250022, People's Republic of China
First published on 19th November 2024
Manganese-based organic metal halides, due to their excellent optoelectronic properties and stable chemical nature, have been widely used in optoelectronic devices and sensors. In this study, through rational modulation of halogens, three novel zero-dimensional organic manganese halides [(C6H8N)2MnX4 (X = Cl, Br, and I)] were obtained by changing the halogen atom via the solvent evaporation method. Due to the d–d electronic transition of Mn2+ in the tetrahedrally coordinated [MnX4]2− polyhedron, all samples exhibit strong green emission. Notably,(C6H8N)2MnBr4 exhibited a remarkable light yield of 18
224 photons per MeV and a low detection limit of 1.9 μGy s−1, which is below the X-ray diagnostic limit and even superior to that of commercial BGO scintillators. Moreover, a 9 cm × 9 cm flexible scintillator film was successfully fabricated by mixing (C6H8N)2MnBr4 crystalline powder with polymethyl methacrylate, and this film manifested superior radiation stability and an X-ray imaging resolution of 12.1 lp mm−1. Notably, the X-ray images captured by the flexible film demonstrate distinguished clarity when adhered perfectly to curved metal sheets. The combination of excellent performance and facile solution processing opens new opportunities for low-cost, high-performance organic metal halide-based large-area flexible scintillators for X-ray detection and imaging.
:
Tl, and CsI
:
Tl, have been successfully applied to X-ray imaging due to their high X-ray absorption capacity.7 However, the high-temperature conditions required for growing these crystals make the production process complex and expensive.8,9 Therefore, research is imperative for achieving low-cost and high-performance scintillators.
In recent years, organic metal halide materials have become a focal point in the research of new-generation scintillator materials. These materials are favored for their cost-effective production, adjustable band gaps, and excellent luminescence properties, presenting vast opportunities in light-emitting diodes and detectors.10–13 Among them, organic lead halide perovskite materials demonstrate suitability for X-ray imaging, benefiting from their high X-ray attenuation ability, and have achieved notable advancements.14,15 However, the high toxicity and poor stability of lead-based perovskites limit their further practical application. With reports of scintillator materials based on the metals of tin (Sn), copper (Cu), antimony (Sb) and manganese (Mn), such as crystals of (C8H17NH3)2SnBr4,16 (4-bzpy)4Cu4I4,17 C50H44P2SbCl5
18,19 and (C38H34P2)MnBr4,20 new environmentally friendly lead-free counterparts have emerged.21 It is noteworthy that organic manganese halides stand out among other materials due to their unique luminous mechanism, high luminous efficiency, facile synthesis, and simple structure design, sparking waves of research enthusiasm.22
Compared with other lead-free halide materials, the luminescence of organic manganese halide scintillators originates from the energy transition emission of the Mn2+ luminescence center.23 It can be excited in the UV or blue region, exhibiting a larger Stokes shift and a high PLQY.24 The luminescence intensity is influenced by the crystal field intensity. Depending on the coordination environment of Mn2+ ions, octahedral Mn2+ halides exhibit red emission, while tetrahedral Mn2+ halides exhibit green emission.25 In addition, their low-dimensional crystal structure endows them with enhanced electron localization, which promotes radiative recombination and results in higher PLQY.22 For example, Jiang et al.24 reported that the C4H12NMnCl3 crystal with Mn2+ octahedral coordination exhibits red emission at 635 nm, with a PLQY as high as 91.8%. This scintillator demonstrates good radiation hardness and can achieve X-ray imaging resolutions of up to 5 lp mm−1. In addition, previously reported crystals such as TEA2MnI4,26 TPP2MnBr4
27 and (C38H34P2)MnBr4,23 which all have tetrahedrally coordinated Mn2+, exhibit green light emission, and possess excellent photophysical properties. Their light yield (LY) can even rival that of commercial scintillators, showcasing immense potential in the field of X-ray imaging.
The photophysical properties of organic metal halides are significantly influenced by various halogen elements.28 Each halogen not only affects the chemical stability of these materials, but also alters their optical characteristics. For instance, the previous literature indicates that owing to the smaller radius and higher electronegativity of chlorine, MAPbCl3 exhibits stronger ionic bond stability in thermal environments compared to MAPbBr3 and MAPbI3.29,30 In contrast, the incorporation of bromine often yields a favorable balance between light absorption and luminescence efficiency,31 demonstrating considerable potential for applications. Additionally, iodine, with its larger atomic radius and atomic number, enhances the adaptability of the crystal structure and increases X-ray attenuation.26 Thus, the careful selection and manipulation of different halogens are crucial for optimizing the optoelectronic properties of organic metal halides, providing valuable insights for the development of new and efficient optoelectronic devices.
For X-ray imaging equipment, such as commercial computed X-ray tomography scanners, the internal X-ray detection component is typically designed in a ring shape to achieve comprehensive scanning of the target object. Unfortunately, most available scintillators are rigid single crystals (e.g., NaI
:
Tl and CsI
:
Tl) that are assembled into rings using multiple scintillators,32,33 resulting in loss of image information at the joints. Therefore, there arises a demand for a large-area flexible film that can adapt to detectors of various shapes and configurations. Additionally, for irregularly shaped imaging objects, a flexible scintillation film can conform to the object's surface, reducing scattering lines and optical crosstalk, consequently improving image quality and spatial resolution.34
Herein, three types of 0D manganese-based organic metal halides [(C6H8N)2MnX4 (X = Cl, Br, and I)] were rationally designed and prepared via solvent evaporation. 2-Methylpyridine was selected as the organic ligand for its ability to effectively integrate into [MnX4]2−, consequently increasing the Mn–Mn distance. Furthermore, its strong interactions with halides can enhance structural rigidity and reduce non-radiative transitions from thermal vibrations.24,25,35 Based on the d–d transition of the [MnX4]2− tetrahedron, the three compounds exhibit efficient green luminescence, with (C6H8N)2MnBr4 achieving a high PLQY of 64.7% that significantly surpasses the 17.18% of (C6H8N)2MnCl4 and the 15.31% of (C6H8N)2MnI4. Furthermore, the single crystals of (C6H8N)2MnBr4 display extraordinary scintillation properties, exhibiting an LY of 18
224 photons per MeV and a low detection limit of 1.9 μGy s−1. By mixing the crystal powder with PMMA, a large-area flexible scintillation film was successfully fabricated with size of 9 cm × 9 cm. This film was utilized for imaging objects, achieving a spatial resolution of 12.1 lp mm−1. Additionally, the images of bent metal sheets further validated the superior performance of the flexible film in X-ray imaging. The film demonstrates its capability to conform to various complex and irregular surfaces, highlighting its vast range of application prospects in medical, industrial, security, and other fields.
The powder X-ray diffraction (PXRD) results of the three crystals align with the SCXRD simulation results, confirming that the synthesized SCs possess a high-purity phase (Fig. 2(a)). The Fourier transform infrared (FTIR) spectra show notable peaks, including those for C–N stretching at 1099 cm−1 and N–H vibration at 1565 cm−1. These observations determine the presence of organic molecules (Fig. 2(b)). Additionally, the high-resolution X-ray photoelectron spectroscopy (XPS) spectra also demonstrate the presence of C, N, Mn, and X (X = Cl, Br, and I) elements in the three crystals (Fig. 2(c)). As shown in Fig. S3–S5,† there are two distinct Mn 2p characteristic peaks in the (C6H8N)2MnX4 (X = Cl, Br, and I) crystals, corresponding to Mn 2p3/2 and Mn 2p1/2 respectively. These peaks are observed at 641.7 and 653.5 eV for (C6H8N)2MnCl4, 641.8 and 653.8 eV for (C6H8N)2MnBr4, and 640.6 and 652.6 eV for (C6H8N)2MnI4. Additionally, the two peaks at 198.1 and 199.7 eV in Fig. S3(b)† can be attributed to Cl 2p3/2 and Cl 2p1/2 in the (C6H8N)2MnCl4 crystals. The peaks at 68.4 and 69.4 eV correspond to Br 3d5/2 and Br 3d3/2 in the (C6H8N)2MnBr4 crystals (Fig. S4(b)†), while the peaks at 617.6 and 629 eV are assigned to I 3d2/5 and I 3d3/2 in the (C6H8N)2MnI4 crystals (Fig. S5(b)†), respectively. The results of scanning electron microscopy (SEM) and energy dispersive spectroscopy (EDS) of the (C6H8N)2MnX4 (X = Cl, Br, and I) SCs in Fig. S6–S8† further verify that elements such as C, N, Mn and X (X = Cl, Br, and I) are present and evenly distributed in the crystals. Furthermore, the thermal stability of the three halide crystals was evaluated through thermogravimetric analysis (TGA), showing that the (C6H8N)2MnX4 (X = Cl, Br, and I) crystals began to decompose at 156 °C, 185 °C, and 247 °C, respectively (Fig. 2(d)).
![]() | ||
| Fig. 2 (a) Experimental PXRD and simulated PXRD patterns, (b) FTIR spectra, (c) survey XPS spectra and (d) TG curves of (C6H8N)2MnX4 (X = Cl, Br, and I). | ||
The steady-state photoluminescence (PL) and excitation (PLE) spectra reveal that all three organic manganese halide crystals exhibit narrow-band green emission in the visible light range (Fig. 3(a–c)). Notably, the PL and PLE spectra are red-shifted with increasing halogen atomic number. Specifically, the peak emission for the crystals based on Cl−, Br−, and I− are centered at 520, 525, and 548 nm, respectively. The emission wavelength is influenced by several factors, particularly the strength of the ligand field and the splitting of the d-orbitals in a tetrahedral environment. For Mn(II) ions, a weaker ligand field results in smaller d-orbital splitting.38 Consequently, samples incorporating Br− exhibit a slight red shift compared to those with Cl−. Additionally, the covalent bond strength of the Mn–I interaction affects the electronic environment of the Mn ions and their d-orbital splitting. Strong covalent bonds can lower the energy required for electronic transitions, leading to further red shift in the emission wavelength.39 UV-vis absorption spectra are used to study the absorption characteristics of the three manganese halide crystals (Fig. 3(d–f)). The three samples exhibit multiple absorption peaks at 263, 298, 364, 377, 434, 455, and 476 nm, corresponding to the 6A1 → 4A2(F), 6A1 → 4T1(F), 6A1 → 4Tg(4p), 6A1 → 4Eg(4D), 6A1 → 4Ag, 4Eg(4G), 6A1 → 4T2(4G) and 6A1 → 4T1(4G) transitions, respectively. After excitation, the excited electrons relax to the lowest excited state 4T1 through several non-radiative processes. Subsequently, they emit bright light via the 4T1 → 6A1 radiative transition (Fig. S9†). Additionally, the direct band gaps of (C6H8N)2MnX4 (X = Cl, Br, and I) were calculated using the Tauc method, yielding values of 2.50, 2.45, and 2.33 eV, respectively (inset of Fig. 3(d–f)). In the tetrahedral coordinated environment of manganese halides, the weak crystal field strength results in a high separation level of 4T1 → 6A1, contributing to the narrow-band green emission of the crystals.24 Afterwards, we recorded the PL and PLE spectra of the three samples at various excitation and emission wavelengths. The contour plots clearly show that the emission centers of the three samples do not shift (Fig. S10†). This consistency further indicates that the green light emission characteristics of (C6H8N)2MnX4 (X = Cl, Br, and I) arise solely from the tetrahedrally coordinated manganese center and are the result of relaxation from the same excited state. Additionally, the decay curves of time-resolved photoluminescence (TRPL) spectra for the (C6H8N)2MnX4 (X = Cl, Br, and I) crystals were fitted using a single exponential function. The calculated lifetimes are 509.4, 262.6 and 15.8 μs, respectively, showcasing a discernible trend of faster attenuation from Cl to Br to I (Fig. 3(g–i)). This trend is mainly due to the heavy-atom-enhanced reverse intersystem crossing (RISC) process.40,41
![]() | ||
| Fig. 3 (a–c) PLE and PL spectra, (d–f) UV-vis absorption spectra, and (g–i) PL decay curves of (C6H8N)2MnX4 (X = Cl, Br, and I). | ||
To evaluate the photophysical properties of the title samples, their PLQYs were measured as 17.18%, 64.7%, and 15.31%, respectively (Fig. S11†). Studies have shown that the Mn–Mn distance significantly influences the luminescence properties of the crystals. Generally, a larger Mn–Mn distance can reduce non-radiative transitions between manganese ions, thereby improving the excitation and recombination efficiency of the photo-generated carriers, which can effectively enhance the PLQY.42,43 Moreover, lower symmetry can accelerate PL decay, which is unfavorable for enhancing the PLQY.43 Consequently, (C6H8N)2MnBr4 exhibits a higher PLQY compared to that of (C6H8N)2MnCl4, which may arise from a longer Mn–Mn distance and a higher symmetry as suggested from the single-crystal structure data. It is also noteworthy that the lower PLQY of (C6H8N)2MnI4 may be ascribed to the significant structural distortion of the inorganic unit, as well as the enhanced spin–orbit coupling effects and strong self-absorption.28,42,43 Specifically, the larger atomic radius and lower electronegativity may lead to a stronger spin–orbit coupling effect, increasing the probability of non-radiative recombination. On the other hand, the substantial overlap between the absorption and emission peaks can enhance the self-absorption of the crystal.28,42 This results in some of the excitation light being absorbed by the material rather than being effectively converted to light emission, ultimately leading to a lower PLQY.
Using density functional theory (DFT), we performed the calculation and investigation of the electronic properties of (C6H8N)2MnBr4, as illustrated in Fig. 4(a). Our analysis reveals that the valence band maximum (VBM) and the conduction band minimum (CBM) are situated at the same high-symmetry point, resulting in a calculated direct bandgap of 2.4 eV, which closely aligns with the experimental values. Further examination indicates that both the VBM and CBM exhibit a significant degree of flatness, suggesting that electrons are in a highly localized state.44 Through partial density of states (PDOS) analysis, we observed that the VBM is primarily composed of Br-p and Mn-d orbitals, while the CBM is predominantly contributed to by the organic components (Fig. 4(b)). According to previous reports, this result indicates that organic molecules not only play an isolating role within the structure, but also transfer energy to the luminescent [MnX4]2− units during the luminescence process.28,45,46
![]() | ||
| Fig. 4 (a) Electronic energy band structures of (C6H8N)2MnBr4. (b) Orbital projection of the partial density of states of (C6H8N)2MnBr4. | ||
To further assess the scintillation performance of the (C6H8N)2MnX4 (X = Cl, Br, and I) crystals, we first calculated the relationship between the X-ray absorption coefficient and the photon energy for the three different samples based on the XCOM database. The results are shown in Fig. 5(a). It is evident that the X-ray absorption coefficients of (C6H8N)2MnX4 (X = Cl, Br, and I) exhibit a progressive increase with the halogen atomic number. This is attributed to their high atomic number, which can effectively interact with X-rays, leading to significant increases in their absorption cross-section for X-rays.47 Subsequently, we plotted the thickness-dependent attenuation efficiency at a photon energy of 22 keV (the average photon energy under 50 kV X-ray irradiation). It is noteworthy that despite the absorption coefficient and attenuation efficiency of these three samples being lower than those of commercial BGO scintillators, the absorption coefficient of (C6H8N)2MnI4 and (C6H8N)2MnBr4 is significantly superior to those of previously reported scintillators such as (C38H34P2)MnBr4,23 (HTPP2)MnBr4,48 C4H12NMnCl3,24 and (ETP)2MnBr4
7 in the range of 40–120 keV. As shown in Fig. 5(b), the attenuation efficiencies of the 1 mm-thick (C6H8N)2MnX4 (X = Cl, Br, and I) scintillator materials for 22 keV X-ray photons are 51.5%, 98.8% and 96.2%, respectively. Notably, at this incident photon energy, the attenuation efficiency of the Br−-based sample is significantly higher than that of the I−-based sample. This phenomenon can be attributed to the fact that the incident photon energy is located near the K absorption edge of (C6H8N)2MnBr4. In this energy range, the photoelectric effect is significantly enhanced, resulting in a notable increase in the absorption capacity of X-rays as they pass through the material.49
The light yield (LY) is another critical parameter for evaluating scintillator performance, as it affects the sensitivity and detection limits of the X-ray detector.1 For comparison, a commercial BGO scintillator serves as a standard reference, along with a mini X-ray generator equipped with a tungsten target and a spectrometer to obtain radioluminescence (RL) spectra. When the scintillator material interacts with incident high-energy X-rays, numerous electron–hole pairs are generated. These electron–hole pairs are then transferred to the tetrahedrally coordinated manganese luminescent centers. Ultimately, the excited Mn2+ ions relax from the 4T1 state to the 6A1 state, emitting scintillation light.26 As depicted in Fig. 5(c), the RL spectra of (C6H8N)2MnX4 (X = Cl, Br, and I) share similar characteristics with their PL spectra, indicating identical radiative recombination pathways for both forms of excitation. At equivalent dose rates, the RL intensity of (C6H8N)2MnBr4 dramatically surpasses that of (C6H8N)2MnCl4, (C6H8N)2MnI4 and the commercial BGO scintillator. The BGO reference sample has an LY of 10
600 photons per MeV.50 In comparison, the calculated LY values for (C6H8N)2MnX4 (X = Cl, Br, and I) are 3009, 18
224 and 3770 photons per MeV, respectively. Additionally, we collected RL spectra of the samples at different dose rates (Fig. 5(d) and Fig. S12†). The (C6H8N)2MnX4 (X = Cl, Br, and I) crystals exhibit perfectly linear responses in the dose rate range of 0.96–5.92 mGy s−1. By calculation, the detection limit for (C6H8N)2MnX4 (X = Cl, Br, and I) is 23.2, 1.9, and 6.1 μGy s−1, respectively. Notably, under the same measurement conditions, the detection limit of (C6H8N)2MnBr4 is only a seventh of that of the BGO scintillator and is under the critical threshold for the X-ray diagnostic limit (5.5 μGy s−1).51–53
To further investigate the temperature-dependent radiation stability of the three samples, their RL spectra were recorded over the temperature range of 300–380 K (Fig. S13†). The results indicate that under the same conditions, (C6H8N)2MnBr4 exhibits lower attenuation and higher stability of RL intensity. At 380 K, its strength reduces to 59% of the initial value, which is significantly better than the 12% recorded for (C6H8N)2MnCl4 and the 24% for (C6H8N)2MnI4. This phenomenon underscores the fact that (C6H8N)2MnBr4 not only performs excellently at ambient temperature, but also maintains excellent stability and effectiveness in high-temperature environments. The above-mentioned exceptional performance of (C6H8N)2MnBr4 establishes it as a strong contender for use in future efficient and reliable X-ray imaging scintillators.
Considering the high X-ray attenuation efficiency and LY of (C6H8N)2MnBr4 crystals, we further investigated their potential as high-resolution X-ray imaging scintillators. Powdered (C6H8N)2MnBr4 crystals were mixed with PMMA to prepare large-area flexible scintillation screens, as shown in Fig. 6(a) and Fig. S14.† The (C6H8N)2MnBr4@PMMA film was approximately 9 cm × 9 cm with a thickness of 79 μm, emitting uniform green light under 365 nm UV irradiation and exhibiting high transparency under daylight, with clearly discernible letters beneath the film (Fig. 6(a) and Fig. S15†). Additionally, the film demonstrated remarkable flexibility, capable of withstanding multiple folds (Fig. 6(a)). Moreover, the SEM-EDS images confirm the uniform distribution of elements within the film (Fig. S16†). Furthermore, the X-ray-induced RL characteristics of the film and crystal remained consistent, indicating no alteration in the luminescent centers (Fig. S17†). To further assess the stability of the scintillating film, we performed continuous irradiation at a dose rate of 8.99 mGy s−1. As illustrated in Fig. 6(b), when the cumulative dose reached 9 Gy, the RL intensity showed no significant changes, indicating that the film possesses excellent radiation stability.
The X-ray imaging quality of (C6H8N)2MnBr4@PMMA flexible films has been systematically studied using a self-constructed imaging system, as depicted in Fig. S18.† Due to the varying X-ray absorption capacities of materials with different densities, the differential RL intensity of the scintillator films captures these variations, allowing optical cameras to obtain internal images of objects.54 At a tube voltage of 50 kV, the internal information of a microchip, a spring-loaded pen, and an earphone was clearly visible, which was imperceptible to the naked eye in daylight (Fig. 6c). To delve deeper into the potential of (C6H8N)2MnBr4 in medical imaging, we selected a shrimp, chicken feet, and a small crab as subjects for experimentation. The X-ray images revealed the internal bones and other tissue structures of these organisms, highlighting the promising application of (C6H8N)2MnBr4@PMMA scintillator films in medical diagnostics.
As shown in Fig. 7(a), the spatial resolution of a standard X-ray image is approximately 12 lp mm−1. By calculating the modulation transfer function (MTF) of the tilting edge X-ray image of the scintillator film (Fig. 7(b)), the spatial resolution was quantified as being 12.1 lp mm−1. This demonstrates its excellent performance in high-resolution imaging applications.
The flexible film can adjust to different shapes and sizes of objects, offering superior fit and coverage. This adaptability makes X-ray imaging of complex structures or irregular surfaces more efficient and precise.55 Moreover, the flexible film can minimize artifacts and distortions that occur when a rigid detector cannot fully conform to the object's surface.56 This enhancement boosts the resolution and clarity of X-ray imaging, resulting in more accurate diagnoses and analytical outcomes. As displayed in Fig. 7(c and d), (C6H8N)2MnBr4@PMMA exhibits a significant difference in X-ray imaging quality between images of a metal sheet in its curved and flat states. When the scintillation film is pressed against the irregular metal sheet, the image becomes clearer and the overall image quality is significantly improved. Therefore, the scintillation film not only enhances the resolution and clarity of the image, but also demonstrates its substantial application potential in flexible X-ray imaging.
224 photons per MeV and a low detection limit of 1.9 μGy s−1, outperforming commercial BGO scintillators and showcasing extensive application prospects. More importantly, the flexible film presented clearer imaging on bent metal sheets in contrast to planar imaging, making it more effective for detecting irregularly shaped objects or covering large areas. This work enriches the excellence of organic manganese halides and demonstrates their latent application prospects in medical imaging, industrial inspection, and security screening.
:
1 were first dissolved in a mixed solution of HCl and CH3OH (1/3, v/v) and the mixture was then heated and stirred at 90 °C to form a transparent precursor solution. Subsequently, the solution was then slowly evaporated on a heating plate at 60 °C in an atmospheric environment. After a few days, light green crystals of (C6H8N)2MnCl4 can be obtained.
:
1 were dissolved in a mixed solution of HBr and CH3OH (1/3, v/v) and the mixture was then heated and stirred at 90 °C to form a transparent precursor solution. Subsequently, the solution was then slowly evaporated on a heating plate at 60 °C in an atmospheric environment. After a few days, green crystals of (C6H8N)2MnBr4 can be obtained.
| AE = [1 − exp(−tρd)] × 100% |
600 photons per MeV) as a reference, we calculated the LYs of (C6H8N)2MnX4 (X = Cl, Br, and I). The experiments were conducted at a tube voltage of 50 kV, with the sample powder compressed to a uniform thickness of approximately 1 mm. The attenuation efficiencies (AE(d)) for (C6H8N)2MnX4 (X = Cl, Br, and I) and BGO can be extracted from Fig. 5(b). The photon counting results (PCmeasured) were obtained by integrating the RL spectrum. Finally, the relative LY was calculated using the following eqn (1) and (2):![]() | (1) |
![]() | (2) |
Footnote |
| † Electronic supplementary information (ESI) available: Single-crystal X-ray diffraction data, XPS spectra, excitation wavelength-dependent PL and PLE spectra, X-ray dose rate-dependent RL spectra, RL spectrum of the scintillation film and X-ray images. CCDC 2270667, 2381207 and 2381212. For ESI and crystallographic data in CIF or other electronic format see DOI: https://doi.org/10.1039/d4qi02522a |
| This journal is © the Partner Organisations 2025 |