Next Article in Journal
Sorption of Polycyclic Aromatic Sulfur Heterocycles (PASH) on Nylon Microplastics at Environmentally Relevant Concentrations
Previous Article in Journal
Enhanced Solubility of Ibuprofen by Complexation with β-Cyclodextrin and Citric Acid
Previous Article in Special Issue
Modulating the ESIPT Mechanism and Luminescence Characteristics of Two Reversible Fluorescent Probes by Solvent Polarity: A Novel Perspective
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Boosting Blue Self-Trapped Exciton Emission in All-Inorganic Zero-Dimensional Metal Halide Cs2ZnCl4 via Zirconium (IV) Doping

1
School of Semiconductors and Physics, North University of China, Taiyuan 030051, China
2
School of Chemistry and Chemical Engineering, Shandong University, Jinan 250100, China
3
Traffic Information Engineering Institute, Guangxi Transport Vocational and Technical College, Nanning 530004, China
*
Authors to whom correspondence should be addressed.
Molecules 2024, 29(7), 1651; https://doi.org/10.3390/molecules29071651
Submission received: 19 March 2024 / Revised: 1 April 2024 / Accepted: 2 April 2024 / Published: 6 April 2024
(This article belongs to the Special Issue Theoretical Study on Luminescent Properties of Organic Materials)

Abstract

:
Low-dimensional metal halides with efficient luminescence properties have received widespread attention recently. However, nontoxic and stable low-dimensional metal halides with efficient blue emission are rarely reported. We used a solvothermal synthesis method to synthesize tetravalent zirconium ion-doped all-inorganic zero-dimensional Cs2ZnCl4 for the first time. Bright blue emission in the range of 370 nm–700 nm with a emission maximum at 456 nm was observed in Zr4+:Cs2ZnCl4 accompanied by a large Stokes shift, which was due to self-trapped excitons (STEs) caused by the lattice vibrations of the twisted structure. Simultaneously, the PLQY of Zr4+:Cs2ZnCl4 achieve an impressive 89.67%, positioning it as a compelling contender for future applications in blue-light technology.

1. Introduction

The broad prospects of solid-state lighting and full-color displays have promoted the rapid development of luminescent materials. Red, green, and blue light sources covering the entire visible light range are required to produce effective white light. The design and manufacture of blue light has proven to be the most challenging of the three sources. For example, blue light-emitting diodes became available two decades after red and green light-emitting diodes were successfully manufactured using epitaxial inorganic technology [1,2].
In the past decade, low-dimensional metal halide luminescent materials have been extensively studied. In particular, zero-dimensional (0D) metal halides have attracted widespread attention due to their high photoluminescence quantum yields (PLQYs) [3]. Various crystal materials including tetrahedral BX4, pyramidal BX5, and octahedral BX6 have been reported in metal halides with a 0D structure [4,5,6]. Green, yellow, and red emission with high PLQY was achieved in 0D metal halides [5,7,8,9,10]. However, blue-emitting metal halides with high efficiency and stability remain challenging. Recently, some lead-based low-dimensional metal halides with efficient blue-light emission have been reported [11,12,13]. Despite the fascinating optical properties of these Pb-based compounds, the toxicity of Pb2+ hinders their further practical applications. Therefore, it is particularly important to design low-toxic or non-toxic low-dimensional metal halide materials with efficient blue emission properties for future applications. At present, Cu(I)-based compounds with high-efficiency blue emission characteristics are the most representative low-toxic halide materials and have received widespread attention from the scientific community [14,15,16,17,18]. The luminescence of Cu(I)-based low-dimensional metal halides comes from the self-trapped excitons formed by its soft lattice structure. The existence of organic macromolecules and large-sized inorganic cations enable the production of a strong spatial confinement effect, resulting in photogenerated excitons. These photogenerated excitons are strongly bound within the inorganic polyhedron which then emits photons outward through radiative recombination, finally producing efficient luminescence. However, the naturally unstable and easily oxidized characteristics of Cu(I) are still a problem that needs to be solved urgently. Xia et al. reported a zero-dimensional Sn(IV)-based organic–inorganic hybrid meatal halide (C6N2H16Cl)2SnCl6 [19]. This material exhibited broadband blue-light emission, which came from the self-trapped excitons formed inside the crystal. At the same time, (C6N2H16Cl)2SnCl6 also had ultra-high stability. The PL intensity of (C6N2H16Cl)2SnCl6 remained stable for more than three months. Its PL intensity remained at about 50% of the initial intensity when the temperature rose to 450 K. However, the emission efficiency of (C6N2H16Cl)2SnCl6 was only 8.1%, which is too low for large-scale commercial applications. Through theoretical calculations, they confirmed that the lower PLQY of (C6N2H16Cl)2SnCl6 was caused by the disappearance of the lone-pair electrons on the Sn4+ cations, which led to the smallest distortion and finally led to the low PLQY. In addition to Sn(IV)-based compounds, In(III)-based compounds have also attracted much attention recently. Saparov et al. reported a new zero-dimensional organic–inorganic hybrid indium bromide RInBr4 [R = trimethyl(4-stilbenyl)methylammonium cation] [20]. The RInBr4 showed bright-blue emission at 437 nm under the 391 nm UV excitation with Commission Internationale de l’Eclairage color coordinates of (0.19, 0.20) and a high photoluminescence quantum yield of 16.36% at room temperature. In addition, this novel hybrid indium bromide demonstrated significantly improved environmental stability. In addition to indium bromide, low-dimensional organic–inorganic hybrid indium chloride with blue-light emission has also been reported. Yue et al. reported two indium chlorides with blue-light emission: [H2EP]2InCl6·Cl·H2O·C3H6O and [H3AEP]InCl6··H2O (EP = 1-ethylpiperazine, AEP = N-aminoethyl piperazine) [21]. The blue emission of both compounds came from self-trapped excitons generated by their soft lattice structures. However, the PLQY of [H2EP]2InCl6·Cl·H2O·C3H6O and [H3AEP]InCl6··H2O was not high. The highest was only 13.44%, which is far from meeting the needs of practical applications. A series of zero-dimensional In-based halides: [H3AEPz]2InCl9·(H2O)2 (AEPz = N-aminoethylpiperazine), [H2AMPd]2InCl7 [AMPd = 4-(aminomethyl)piperidine], and [H2PhPz]2InCl7·(H2O)2 (PhPz = 1-phenylpiperazine), reported by Lei et al., had a maximum PLQY of no more than 20% [22]. Therefore, In(III)- and Sn(IV)-based materials are still plagued by problems such as low luminous efficiency and poor optical properties, which cannot meet the needs of practical applications [19,20,21,22].
In recent years, low-dimensional Zn(II)-based materials based on Zn with a d10 electronic configuration have gradually attracted the attention of researchers due to their variable crystal structure, low toxicity, and high stability properties. Currently, efficient emission is achieved in the all-inorganic zero-dimensional Zn-based compound Cs-Zn-X (X = Cl, Br, and I) by doping modification with Cu+ ions. Through structural transformation and halogen control, the emission color of Cs-Zn-X can be effectively adjusted from violet to blue [5,15,23,24]. However, these materials still face the problem of poor stability due to the presence of Cu2+ ions. After being left for a period of time, their optical performance will gradually deteriorate due to the oxidation of Cu+ to Cu2+ until they lose optical activity. The key to solving this problem is whether we can find a stable ion as a dopant and incorporate it into the Cs-Zn-X crystal lattice so that it can produce efficient and stable blue emissions. We have noticed that the all-inorganic metal halide double perovskite material Cs2ZrCl6 has received widespread attention recently, and this material can produce blue-light emission under ultraviolet light excitation [25,26]. Research results show that this blue luminescence comes from STEs generated by the strong electron–phonon coupling effect of the [ZrCl6]2- inorganic octahedron in Cs2ZrCl6 [26]. Therefore, Zr4+ ions can be incorporated into host materials as optically active ions to achieve blue-light emission.
Herein, we used nontoxic and stable all-inorganic Cs2ZnCl4 as the host and Zr4+ ions as the dopant to design a zero-dimensional metal halide Zr4+:Cs2ZnCl4 with efficient blue emission. By introducing Zr4+ into non-luminescent Cs2ZnCl4, efficient blue-light emission (λem = 456 nm under 260 nm UV excitation) was achieved with a PLQY as high as 89.67%. Detailed optical characterization including photoluminescence (PL) decay lifetime and temperature-dependent PL spectra demonstrated that the bright blue emission in Zr4+:Cs2ZnCl4 originated from self-trapped exciton (STE) emission. We believe that this work can provide a reference for the future design and synthesis of low-toxic metal halide materials with efficient blue emission properties.

2. Results and Discussion

2.1. Structure and Morphology

Pure Cs2ZnCl4 and Zr4+-doped Cs2ZnCl4 crystal samples were prepared by the solvothermal method using cesium chloride, zinc chloride, and zirconium chloride. Crystallographic data indicated that Cs2ZnCl4 belonged to the orthorhombic Pnma space group and had a typical 0D structure. In each [ZnCl4]2− tetrahedron, one Zn2+ ion occupied the central position and four Cl ions were connected to it to form a highly symmetrical tetrahedral geometry, as shown in Figure 1a. Each [ZnCl4]2− tetrahedron was distributed in isolation and the distance between two adjacent tetrahedrons was 6.6 Å (calculated based on the Zn2+ ion distance). The tetrahedrons exhibited a typical 0D structure at the molecular level. Cs+ ions served as a supporting framework and charge balance in the structure. As shown in the upper part of Figure 1b, the bond length of the Zn-Cl bond was between 2.271 Å and 2.297 Å in each [ZnCl4]2− tetrahedron. The bond length range of the Zn-Cl bond in the [ZnCl4]2− tetrahedron changed to 2.274 Å~2.323 Å after Zr4+ ions were introduced into the lattice of Cs2ZnCl4. Compared with undoped Cs2ZnCl4, the introduction of Zr4+ ions significantly increased the bond length of the Zn-Cl bond, which showed that the tetrahedral distortion in the doped system was greater than the undoped system. In addition, the bond angles of the four Cl-Zn-Cl also changed significantly after Zr4+ doping. In the undoped system, the bond angles of Cl1-Zn-Cl2, Cl2-Zn-Cl3, Cl3-Zn-Cl4, and Cl1-Zn-Cl4 were 106.479°, 109.046°, 109.636°, and 115.361°, respectively. After the introduction of Zr ions, the Cl1-Zn-Cl2 and Cl1-Zn-Cl4 bond angles decreased, while the Cl2-Zn-Cl3 and Cl3-Zn-Cl4 bond angles increased, as shown in the lower part of Figure 1b. The significant structural differences between the undoped system and the doped system also provide evidence that Zr4+ ions replaced a small part of the Zn2+ ions lattice. At the same time, this significant difference also showed the stronger distortion of the crystal lattice after Zr4+ doping, and easily generated the strong electron–phonon coupling effects under excitation, which is an intrinsic condition for efficient light emission [27,28,29].
The powder x-ray diffraction (PXRD) patterns of Cs2ZnCl4 crystal samples at different Zr2+ doping concentrations are shown in Figure 1c. Each pattern was in good agreement with the standard diffraction pattern of Cs2ZnCl4 (ISCD-6062) and no redundant impurity diffraction peaks appeared. This result shows that all samples were successfully synthesized and no impurity phase was produced. The uniformity of these patterns also confirms that the doping of Zr4+ did not change the original crystal structure of the Cs2ZnCl4 host material. At the same time, the diffraction pattern of the sample showed a tendency to move to a lower angle as the Zr4+ feed ratio continued to increase, which confirms the successful doping of Zr4+ ions. Because larger Zr2+ ions (ionic radius: 0.6 Å for Zn2+ and 0.72 Å for Zr4+) were incorporated into the Cs2ZnCl4 lattice, the lattice expanded, leading to the movement of the XRD diffraction peak to a lower angle [30,31]. The elemental composition of 20%Zr4+:Cs2ZnCl4 was strictly characterized qualitatively and quantitatively. Energy spectroscopy (EDS) elemental mapping of the Zr4+:Cs2ZnCl4 crystal showed that Cs, Zn, Cl, and Zr elements were uniformly distributed (Figure S1a), indicating that Zr4+ ions were uniformly doped in the Cs2ZnCl4 lattice, with a high phase purity. The measurement results showed that the atomic ratio of Cs:Zn:Zr:Cl was 33.55:11.26:0.39:54.81 while the feed ratio of Zr4+ ions was 20% (Figure S1b). The results of the elemental analysis showed that when the feed ratio of the Zr4+ ions in the experiment was 20%, the actual atomic percentage entering the Cs2ZnCl4 lattice was only 3.35%. This may have been due to the larger radius and higher valence of Zr4+ ions, which could not easily replace the lattice position of Zn2+.

2.2. Optical Properties

In order to study the optical properties of Zr4+:Cs2ZnCl4, the absorption spectra of Cs2ZnCl4 at different Zr4+ doping ratios were first collected, as shown in Figure 2a. The absorption spectrum of the undoped sample showed that absorption intensity increased slowly as the wavelength decreased and no obvious absorption band appeared in the range from 250 nm to 600 nm. With the introduction of Zr4+ ions, two new absorption bands appeared at 300 nm and 350 nm. The intensity of these two absorption bands increased with the increase in the Zr4+ ion doping ratio. There is no doubt that the optical transition process corresponding to these two absorption bands was related to the exciton characteristics generated by structural distortion, which originated from the incorporation of Zr4+ ions [32,33,34].
Under the excitation of 260 nm UV light, Zr4+:Cs2ZnCl4 exhibited bright broadband blue emission in the range of 370 nm to 700 nm with a PL peak position at 456 nm, Stokes shift of 197 nm, and a full-width half maximum (FWHM) of 119.86 nm. The CIE coordinates of Zr4+:Cs2ZnCl4 and the optical photograph of the sample under 254 nm UV excitation are shown in Figure S2a. The PL intensity changed significantly due to the concentration quenching effect, showing an increase first and then a downward trend with the increase in the feeding ratios of Zr4+ (Figure S2b). When the Zr4+ feeding ratio was 20%, the PL intensity reached the maximum. In order to investigate the potential photophysical mechanism of these Zr4+ doped samples, the PLE spectra of Zr4+:Cs2ZnCl4 at different emission wavelengths and the PL spectra at different excitation wavelengths were collected. Figure S3a shows that the PLE spectra at different emission wavelengths all had similar profiles, which excluded the possibility that the blue emission came from other defects or impurities in the crystal. Figure S3b shows that the PL spectra of the samples obtained with different excitation wavelengths also had the same shape, indicating that this blue emission came from the same excited state relaxation.
Figure 2c shows the PL lifetime decay curve of Cs2ZnCl4 under different Zr4+ doping ratios. The lifetime curves of all samples could be fitted using single exponential decay. All these samples exhibited decay lifetimes in the microsecond range. PL decay lifetime in the order of microseconds is a typical feature of STE emission, and the lifetime of this emission is usually longer than the lifetime of band-edge free excitons [27,35]. As the Zr4+ doping amount gradually increased, the PL lifetime of these samples showed a trend of first increasing and then decreasing, with the longest lifetime at 20% doping amount, which was similar to the change trend of the PL spectrum (Figure S2b). The decrease in lifetime at high doping ratios was due to the concentration quenching effect of high concentrations of Zr4+ ions because high concentrations of Zr4+ ions will generate more non-radiative recombination paths to quench the emission. Characteristics such as broadband emission and long lifetime in the order of microseconds indicated that the bright blue emission in Zr4+:Cs2ZnCl4 came from STE emission caused by strong electron–phonon coupling inside the crystal lattice. This strong electron–phonon coupling effect exists widely in “soft” lattice materials with less structural rigidity [35,36]. The strong electron–phonon coupling effect will cause elastic deformation of the lattice in this “soft” lattice structure, thereby binding the photogenerated excitons to certain specific positions on the lattice to form STEs. This state has a lower energy level than other excited states and, therefore, can exist stably. These excitons are bound to specific positions in the crystal lattice, and can only radiate photons outward through radiative recombination to produce emission. In addition, the sample obtained a high PLQY of 89.67% under 260 nm ultraviolet light excitation when the Zr4+ ion doping amount was 20%. This indicates that Zr4+:Cs2ZnCl4 has certain applications in the field of solid-state lighting and display (Figure S4).
In order to explore the electron–phonon coupling effect inside the sample, a 633 nm continuous laser was used as the excitation source to collect the Raman spectra of Cs2ZnCl4 under different Zr2+ doping ratios, as shown in Figure 2d. In all samples, six Raman peaks could be observed in the low wavenumber region (<500 cm−1). These were located at 69 cm−1, 106 cm−1, 121 cm−1, 134 cm−1, 282 cm−1, and 291 cm−1, respectively. In addition, new Raman peaks appeared near 154 cm−1 and 318 cm−1 when the Zr4+ feed ratios were 20% and 40%, which may have been caused by high-concentration Zr4+ doping. It is worth noting that the Raman peak at 282 cm−1 was the double frequency mode of the 134 cm−1 Raman peak. This double frequency mode indicated that there was a strong electron–phonon coupling in Zr4+:Cs2ZnCl4, which was very conducive to the formation of STE [37]. This result further proves that the broadband blue emission of Zr4+:Cs2ZnCl4 was a typical STE emission, which came from the strong electron–phonon coupling effect caused by the structure distortion.
To further explore the intrinsic photophysical properties in Zr4+:Cs2ZnCl4, the temperature-dependent PL spectrum of Zr4+:Cs2ZnCl4 was measured. Figure 3a shows the map of the PL intensity of Zr4+:Cs2ZnCl4 plotted with temperature T and emission wavelength λ at the temperature range of 80–360 K under 260 nm UV excitation. The color scale represents the PL intensity, and the color change from blue to red represents the enhancement of the PL intensity from the minimum to the maximum at the temperature range of 80 K–300 K. Zr4+:Cs2ZnCl4 exhibited a single emission band (λem = 456 nm) at all temperatures and additional emission bands appeared with temperature changes. In the temperature range of 80 K–150 K, the PL intensity of Zr4+:Cs4ZnCl4 was too weak to show the obvious emission signal. When the temperature was higher than 150 K, the PL intensity of Zr4+:Cs4ZnCl4 began to increase significantly with the increase in temperature and was strongest at 300 K. As the temperature increased further, PL intensity of Zr4+:Cs4ZnCl4 weakened. According to previous research, the formation of STEs requires the participation of a certain intensity of lattice vibration [27,32,35]. In Zr4+:Cs2ZnCl4, the lattice vibration intensity required to generate STEs was stronger. The number of STEs formed was limited and effective emission could not be produced, due to the lattice vibration being significantly suppressed at low temperatures. The enhanced lattice vibration intensity promoted the formation of STEs when the temperature gradually increased, resulting in enhanced PL intensity. Finally, the promotion effect of lattice vibration on emission reached the maximum at 300 K, and the PL intensity was the highest. With further temperature increase, the strong lattice vibration will no longer promote emission but will produce a quenching effect. Therefore, the PL intensity of Zr4+:Cs2ZnCl4 began to decrease when the temperature was higher than 300 K.
In addition, several key physical parameters, such as the Huang–Rhys factor (S), the exciton binding energy (Eb), and the electron–optical-phonon coupling energy (Γop) were obtained by fitting the temperature-dependent PL spectrum. These physical parameters help to understand the underlying photophysical mechanism in Zr4+:Cs2ZnCl4. First, the exciton binding energy of the Zr4+:Cs2ZnCl4 was fitted using the following formula [38]:
I ( T )   = I 0 1 + A e E b k b T
where I0 is the PL intensity at 0 K, I(T) is the PL intensity at temperature T K, kb is the Boltzmann constant, and A is a constant. The fitting results, as given in Figure 3b, showed that the Eb of the Zr4+:Cs2ZnCl4 was 481.98 meV. Such a large exciton binding energy indicated that these STEs required high energy to generate, which was consistent with the optical behavior at different temperatures. At the same time, the larger the exciton binding energy, the more stable the photogenerated excitons and the higher the dissociation temperature, thereby ensuring that the emission had excellent thermal stability and resistance to thermal quenching.
The Huang–Rhys factor (S) of the sample was obtained to investigate the electron–phonon coupling effect by fitting using the following formula [39]:
FWHM T = 2.36 S ω phonon coth ω phonon 2 k b T
where ωphonon is the phonon frequency, T is the temperature, FWHM is the full-width half maximum of the PL spectra, and kb is the Boltzmann constant. The fitting results are shown in Figure 3c. The Huang–Rhys factor S of 40.34 meant that Zr4+:Cs2ZnCl4 had strong electron–phonon coupling. At the same time, the phonon energy ℏωphonon obtained by fitting was 40.89 meV, which was very close to the vibration energy of the 291 cm−1 Raman peak in the Raman spectrum (~36.08 meV). This indicates that the 291 cm−1 Raman peak mainly participated in the formation of STEs in Zr4+:Cs2ZnCl4. The large S value, broadband emission, and large Stokes shift (197 nm) indicated that the blue emission of Zr4+:Cs2ZnCl4 originated from self-trapped excitons generated by strong Jahn–Teller distortion. The Toyokawa equation was used to fit the change in PL spectrum FWHM with temperature to further investigate the electron–phonon coupling effect in the Zr4+:Cs2ZnCl4 [40]:
Γ ( T )   = Γ 0 + Γ op e ω o p k b T 1
where Γ0 is the FWHM of PL spectra at 0 K, Γop is the electron–optical-phonon coupling energy, ℏωop is the energy of long optical phono, and kb is the Boltzmann constant. ℏωop was obtained by fitting the Huang–Rhys factor S, which was 40.89. As shown in Figure 3d, the fitted Γ0 and Γop were 611.07 meV and 377.34 meV, respectively. Such a large electron–optical-phonon coupling energy Γ0 indicates that there was a strong electron–phonon coupling in the Zr4+:Cs2ZnCl4 structure, which had a crucial impact on the generation of STEs.

2.3. Theoretical Calculation and Photophysical Model

In order to explore the electronic structure of Zr4+:Cs2ZnCl4 and the intrinsic physical mechanism of efficient blue emission, the bandgap structure and density of state (DOS) of Cs2ZnCl4 and Zr4+:Cs2ZnCl4 were calculated using first principles. Figure 4a,c show the bandgap structures of Cs2ZnCl4 and Zr4+:Cs2ZnCl4, respectively. Interestingly, the band edge after doping with Zr4+ ions was flatter than that of pure Cs2ZnCl4, which means that the doping of Zr4+ ions enhanced the quantum confinement effect because the flat conduction band and valence band meant small dispersion and large carrier effective mass, which is a typical feature of a strong quantum confinement effect [41]. The strong quantum confinement effect was very important for the highly localized exciton formation. Once excitons were formed under excitation, they would be confined to isolated inorganic polyhedrons due to the unique quantum confinement effect of the zero-dimensional structure. There was no interaction due to the large distance between adjacent polyhedrals, leading to highly efficient luminescence in Zr4+:Cs2ZnCl4. Figure 4b,d show the DOS of Cs2ZnCl4 and Zr4+:Cs2ZnCl4, respectively. As shown in Figure 4b, the valence band maximum (VBM) of Cs2ZnCl4 was mainly composed of Zn-3d and Cl-3p orbitals, while Zn-4s, Cl-3p, and Cl-3s orbitals together constituted the conduction band minimum (CBM). The VBM of Zr4+:Cs2ZnCl4 was mainly composed of Zn-3d and Cl-3p orbitals, while the Cl-3p and Zr-4d orbitals together constituted the CBM of Zr4+:Cs2ZnCl4 (Figure 4d). In addition, we also calculated the 3D and 2D electron distribution maps of VBM and CBM in the two compounds, as shown in Figure S5. Through the 3D and 2D electron distribution diagrams of VBM and CBM in pure ((a)–(d)) Cs2ZnCl4 and ((e)–(h)) Zr4+:Cs2ZnCl4, we can intuitively see that the VBM of Cs2ZnCl4 was mainly composed of Zn and Cl orbitals, while Zn and Cl orbitals together constituted the CBM. The VBM of Zr4+:Cs2ZnCl4 was mainly composed of Zn and Cl orbitals, while the Zr orbitals constituted the CBM of Zr4+:Cs2ZnCl4. The DOS of Zr4+:Cs2ZnCl4 showed that its bright blue emission was related to the doping of Zr4+ ions. Combining the PL spectrum and PL decay lifetime, it can be further confirmed that this blue emission originated from self-trapped excitons generated by lattice distortion caused by Zr4+ ion doping. The ultra-high PLQY of close to 100% in Zr4+:Cs2ZnCl4 could be assigned to the highly localized exciton distribution in the structure, which benefited from the strong quantum confinement effect unique to the zero-dimensional structure.
As shown in Figure 5, ground state electrons absorbed energy and transited to the excited state under UV excitation. Due to the soft lattice properties of Zr4+:Cs2ZnCl4, the lattice deformed rapidly after being excited and bound the excited electrons to specific sites in the lattice. The bound electrons then radiated photons outward through radiative recombination and produced bright blue-light emission.

3. Conclusions

In summary, in this work, we report an all-inorganic zero-dimensional metal halide Zr4+:Cs2ZnCl4 with efficient broadband blue emission. Under the excitation of 260 nm UV light, Zr4+:Cs2ZnCl4 exhibited broadband blue emission at 457 nm, with a PLQY of up to 89.67%. Combining the room- and low-temperature PL spectra with the PL decay lifetime, confirmed that the broadband blue emission in Zr4+:Cs2ZnCl4 came from self-trapped excitons generated by structural distortion after Zr4+ ion doping. First-principles calculation results showed that the efficient emission came from the strong quantum confinement effect generated by the unique zero-dimensional structure of Cs2ZnCl4. We believe this work can provide a reference for the future design and synthesis of low-dimensional metal halide materials with high luminescence efficiency.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules29071651/s1, Experimental and computational details and supplementary figures (PDF). Ref. [42] is cited in Supplementary Materials.

Author Contributions

Conceptualization, Y.T. and C.P.; methodology, Y.T.; software, Q.W.; Resources, L.D.; writing—original draft preparation, Y.T. and C.P.; writing—review and editing, Y.T. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the National Natural Science Foundation of China (Grant No. 62175219), the Fundamental Research Program of Shanxi Province (Grant No. 20210302124397) and the Open Research Fund of the State Key Laboratory of Dynamic Testing (Technology Grant No. 2022-SYSJJ-01.

Data Availability Statement

Data are contained within the article and Supplementary Materials.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Ponce, F.; Bour, D. Nitride-Based Semiconductors for Blue and Green Light-Emitting Devices. Nature 1997, 386, 351–359. [Google Scholar] [CrossRef]
  2. Fasol, G.; Nakamura, S. The Blue Laser Diode; Springer-Verlag: Berlin/Heidelberg, Germany, 1997. [Google Scholar]
  3. Zhou, C.; Lin, H.; He, Q.; Xu, L.; Worku, M.; Chaaban, M.; Lee, S.; Shi, X.; Du, M.-H.; Ma, B. Low Dimensional Metal Halide Perovskites and Hybrids. Mater. Sci. Eng. R 2019, 137, 38–65. [Google Scholar] [CrossRef]
  4. Jiang, T.; Ma, W.; Zhang, H.; Tian, Y.; Lin, G.; Xiao, W.; Yu, X.; Qiu, J.; Xu, X.; Yang, Y. Highly Efficient and Tunable emission of Lead-Free Manganese Halides toward White Light-Emitting Diode and X-ray Scintillation Applications. Adv. Funct. Mater. 2021, 31, 2009973. [Google Scholar] [CrossRef]
  5. Han, P.; Luo, C.; Yang, S.; Yang, Y.; Deng, W.; Han, K. All-Inorganic Lead-Free 0D Perovskites by A Doping Strategy to Achieve A PLQY Boost from <2% to 90%. Angew. Chem. Int. Ed. 2020, 132, 12809–12813. [Google Scholar]
  6. Li, Z.; Li, Y.; Liang, P.; Zhou, T.; Wang, L.; Xie, R.-J. Dual-Band Luminescent Lead-Free Antimony Chloride Halides with Near-Unity Photoluminescence Quantum Efficiency. Chem. Mater. 2019, 31, 9363–9371. [Google Scholar] [CrossRef]
  7. Zhou, C.; Lin, H.; Tian, Y.; Yuan, Z.; Clark, R.; Chen, B.; van de Burgt, L.J.; Wang, J.C.; Zhou, Y.; Hanson, K. Luminescent Zero-Dimensional Organic Metal Halide Hybrids with Near-Unity Quantum Efficiency. Chem. Sci. 2018, 9, 586–593. [Google Scholar] [CrossRef] [PubMed]
  8. Zhou, C.; Lin, H.; Shi, H.; Tian, Y.; Pak, C.; Shatruk, M.; Zhou, Y.; Djurovich, P.; Du, M.H.; Ma, B. A Zero-Dimensional Organic Seesaw-Shaped Tin Bromide with Highly Efficient Strongly Stokes-Shifted Deep-Red Emission. Angew. Chem. Int. Ed. 2018, 130, 1033–1036. [Google Scholar] [CrossRef]
  9. Zhou, C.; Worku, M.; Neu, J.; Lin, H.; Tian, Y.; Lee, S.; Zhou, Y.; Han, D.; Chen, S.; Hao, A. Facile Preparation of Light Emitting Organic Metal Halide Crystals with Near-Unity Quantum Efficiency. Chem. Mater. 2018, 30, 2374–2378. [Google Scholar] [CrossRef]
  10. Zhou, L.; Liao, J.F.; Huang, Z.G.; Wei, J.H.; Wang, X.D.; Li, W.G.; Chen, H.Y.; Kuang, D.B.; Su, C.Y. A Highly Red-Emissive Lead-Free Indium-Based Perovskite Single Crystal for Sensitive Water Detection. Angew. Chem. Int. Ed. 2019, 131, 5331–5335. [Google Scholar] [CrossRef]
  11. Morad, V.; Shynkarenko, Y.; Yakunin, S.; Brumberg, A.; Schaller, R.D.; Kovalenko, M.V. Disphenoidal Zero-Dimensional Lead, Tin, and Germanium Halides: Highly Emissive Singlet and Triplet Self-Trapped Excitons and X-ray Scintillation. J. Am. Chem. Soc. 2019, 141, 9764–9768. [Google Scholar] [CrossRef]
  12. Lin, H.; Zhou, C.; Chaaban, M.; Xu, L.-J.; Zhou, Y.; Neu, J.; Worku, M.; Berkwits, E.; He, Q.; Lee, S. Bulk Assembly of Zero-Dimensional Organic Lead Bromide Hybrid with Efficient Blue Emission. ACS Mater. Lett. 2019, 1, 594–598. [Google Scholar] [CrossRef]
  13. Zhou, C.; Lin, H.; Worku, M.; Neu, J.; Zhou, Y.; Tian, Y.; Lee, S.; Djurovich, P.; Siegrist, T.; Ma, B. Blue Emitting Single Crystalline Assembly of Metal Halide Clusters. J. Am. Chem. Soc. 2018, 140, 13181–13184. [Google Scholar] [CrossRef]
  14. Jun, T.; Sim, K.; Iimura, S.; Sasase, M.; Kamioka, H.; Kim, J.; Hosono, H. Lead-Free Highly Efficient Blue-Emitting Cs3Cu2I5 with 0D Electronic Structure. Adv. Mater. 2018, 30, 1804547. [Google Scholar] [CrossRef] [PubMed]
  15. Huang, J.; Su, B.; Song, E.; Molokeev, M.S.; Xia, Z. Ultra-Broad-Band-Excitable Cu(I)-Based Organometallic Halide with Near-Unity Emission for Light-Emitting Diode Applications. Chem. Mater. 2021, 33, 4382–4389. [Google Scholar] [CrossRef]
  16. Creason, T.D.; Yangui, A.; Roccanova, R.; Strom, A.; Du, M.H.; Saparov, B. Rb2CuX3 (X = Cl, Br): 1D All-Inorganic Copper Halides with Ultrabright Blue Emission and Up-Conversion Photoluminescence. Adv. Opt. Mater. 2020, 8, 1901338. [Google Scholar] [CrossRef]
  17. Liu, X.; Yuan, F.; Zhu, C.; Li, J.; Lv, X.; Xing, G.; Wei, Q.; Wang, G.; Dai, J.; Dong, H. Near-Unity Blue Lminance from Lad-Free Copper Halides for Light-Emitting Diodes. Nano Energy 2022, 91, 106664. [Google Scholar] [CrossRef]
  18. Li, H.; Lv, Y.; Zhou, Z.; Tong, H.; Liu, W.; Ouyang, G. Coordinated Anionic Inorganic Module—An Efficient Approach towards Highly Efficient Blue-Emitting Copper Halide Ionic Hybrid Structures. Angew. Chem. Int. Ed. 2022, 61, e202115225. [Google Scholar] [CrossRef] [PubMed]
  19. Song, G.; Li, M.; Yang, Y.; Liang, F.; Huang, Q.; Liu, X.; Gong, P.; Xia, Z.; Lin, Z. Lead-Free Tin (IV)-Based Organic–Inorganic Metal Halide Hybrids with Excellent Stability and Blue-Broadband Emission. J. Phys. Chem. Lett. 2020, 11, 1808–1813. [Google Scholar] [CrossRef] [PubMed]
  20. Fattal, H.; Creason, T.D.; Delzer, C.J.; Yangui, A.; Hayward, J.P.; Ross, B.J.; Du, M.-H.; Glatzhofer, D.T.; Saparov, B. Zero-Dimensional Hybrid Organic–Inorganic Indium Bromide with Blue Emission. Inorg. Chem. 2021, 60, 1045–1054. [Google Scholar] [CrossRef]
  21. Ma, Y.-Y.; Fu, H.-Q.; Liu, X.-L.; Sun, Y.-M.; Zhong, Q.-Q.; Xu, W.-J.; Lei, X.-W.; Liu, G.-D.; Yue, C.-Y. Zero-Dimensional Organic–Inorganic Hybrid Indium Chlorides with Intrinsic Blue Light Emissions. Inorg. Chem. 2022, 61, 8977–8981. [Google Scholar] [CrossRef]
  22. Li, D.-Y.; Sun, Y.-M.; Wang, X.-Y.; Wang, N.-N.; Zhang, X.-Y.; Yue, C.-Y.; Lei, X.-W. Zero-Dimensional Hybrid Indium Halides with Efficient and Tunable White-Light Emissions. J. Phys. Chem. Lett. 2022, 13, 6635–6643. [Google Scholar] [CrossRef] [PubMed]
  23. Babu, R.; López-Fernández, I.; Prasanthkumar, S.; Polavarapu, L. All-Inorganic Lead-Free Doped-Metal Halides for Bright Solid-State Emission from Primary Colors to White Light. ACS Appl. Mater. Interfaces 2023, 15, 35206–35215. [Google Scholar] [CrossRef]
  24. Yang, Y.; Wu, J.; Zhou, T.; Wang, Y.; Zheng, J.; Liu, R.; Hou, J.; Li, X.; Wang, L.; Jiang, W. Efficient Emission in Copper-Doped Cs3ZnX5 (X = Cl, I) for White LEDs and X-ray Scintillators. J. Mater. Chem. C 2023, 11, 9850–9859. [Google Scholar] [CrossRef]
  25. Tang, Y.; Deng, M.; Zhou, Z.; Kang, C.; Wang, J.; Liu, Q. Recent Advances in Lead-Free Cs2ZrCl6 Metal Halide Perovskites and Their Derivatives: From Fundamentals to Advanced Applications. Coord. Chem. Rev. 2024, 499, 215490. [Google Scholar] [CrossRef]
  26. He, Y.; Liu, S.; Yao, Z.; Zhao, Q.; Chabera, P.; Zheng, K.; Yang, B.; Pullerits, T.; Chen, J. Nature of Self-Trapped Exciton Emission in Zero-Dimensional Cs2ZrCl6 Perovskite Nanocrystals. J. Phys. Chem. Lett. 2023, 14, 7665–7671. [Google Scholar] [CrossRef] [PubMed]
  27. Dohner, E.R.; Hoke, E.T.; Karunadasa, H.I. Self-Assembly of Broadband White-Light Emitters. J. Am. Chem. Soc. 2014, 136, 1718–1721. [Google Scholar] [CrossRef]
  28. Mao, L.; Wu, Y.; Stoumpos, C.C.; Wasielewski, M.R.; Kanatzidis, M.G. White-Light Emission and Structural Distortion in New Corrugated Two-Dimensional Lead Bromide Perovskites. J. Am. Chem. Soc. 2017, 139, 5210–5215. [Google Scholar] [CrossRef]
  29. Peng, C.; Zhuang, Z.; Yang, H.; Zhang, G.; Fei, H. Ultrastable, Cationic Three-Dimensional Lead Bromide Frameworks that Intrinsically Emit Broadband White-Light. Chem. Sci. 2018, 9, 1627–1633. [Google Scholar] [CrossRef]
  30. Cheng, P.; Feng, L.; Liu, Y.; Zheng, D.; Sang, Y.; Zhao, W.; Yang, Y.; Yang, S.; Wei, D.; Wang, G. Doped Zero-Dimensional Cesium Zinc Halides for High-Efficiency Blue Light Emission. Angew. Chem. Int. Ed. 2020, 59, 21414–21418. [Google Scholar] [CrossRef]
  31. Wang, J.; Yu, Y.; Li, S.; Guo, L.; Wang, E.; Cao, Y. Doping Behavior of Zr4+ Ions in Zr4+-Doped TiO2 Nanoparticles. J. Phys. Chem. C 2013, 117, 27120–27126. [Google Scholar] [CrossRef]
  32. Dohner, E.R.; Jaffe, A.; Bradshaw, L.R.; Karunadasa, H.I. Intrinsic White-Light Emission from Layered Hybrid Perovskites. J. Am. Chem. Soc. 2014, 136, 13154–13157. [Google Scholar] [CrossRef]
  33. Cortecchia, D.; Yin, J.; Bruno, A.; Lo, S.-Z.A.; Gurzadyan, G.G.; Mhaisalkar, S.; Brédas, J.-L.; Soci, C. Polaron Self-Iocalization in White-Light Emitting Hybrid Perovskites. J. Mater. Chem. C 2017, 5, 2771–2780. [Google Scholar] [CrossRef]
  34. Liu, X.; Peng, C.; Zhang, L.; Guo, D.; Pan, Y. Te4+-Doped Zro-Dimensional Cs2ZnCl4 Single Crystals for Broadband Yellow Light Emission. J. Mater. Chem. C 2022, 10, 204–209. [Google Scholar] [CrossRef]
  35. Li, S.; Luo, J.; Liu, J.; Tang, J. Self-Trapped Excitons in All-Inorganic Halide Perovskites: Fundamentals, Status, and Potential Applications. J. Phys. Chem. Lett. 2019, 10, 1999–2007. [Google Scholar] [CrossRef] [PubMed]
  36. Zhou, L.; Liao, J.F.; Huang, Z.G.; Wei, J.H.; Wang, X.D.; Chen, H.Y.; Kuang, D.B. Intrinsic Self-Trapped Emission in 0D Lead-Free (C4H14N2)2In2Br10 Single Crystal. Angew. Chem. Int. Ed. 2019, 131, 15581–15586. [Google Scholar] [CrossRef]
  37. Peng, C.; Wei, Q.; Yao, S.; Meng, X.; Yu, Z.; Peng, H.; Zhong, X.; Zou, B. H2O–NH4+-Induced Emission Modulation in Sb3+-Doped (NH4)2InCl5·H2O. Inorg. Chem. 2022, 61, 12406–12414. [Google Scholar] [CrossRef] [PubMed]
  38. Yang, B.; Yin, L.; Niu, G.; Yuan, J.H.; Xue, K.H.; Tan, Z.; Miao, X.S.; Niu, M.; Du, X.; Song, H. Lead-Free Halide Rb2CuBr3 as Sensitive X-ray Scintillator. Adv. Mater. 2019, 31, 1904711. [Google Scholar] [CrossRef] [PubMed]
  39. Jing, Y.; Liu, Y.; Jiang, X.; Molokeev, M.S.; Lin, Z.; Xia, Z. Sb3+ Dopant and Halogen Substitution Triggered Highly Efficient and Tunable Emission in Lead-Free Metal Halide Single Crystals. Chem. Mater. 2020, 32, 5327–5334. [Google Scholar] [CrossRef]
  40. Gong, X.; Voznyy, O.; Jain, A.; Liu, W.; Sabatini, R.; Piontkowski, Z.; Walters, G.; Bappi, G.; Nokhrin, S.; Bushuyev, O. Electron–Phonon Interaction in Efficient Perovskite Blue Emitters. Nat. Mater. 2018, 17, 550–556. [Google Scholar] [CrossRef]
  41. Li, Z.; Zhang, C.; Li, B.; Lin, C.; Li, Y.; Wang, L.; Xie, R.-J. Large-Scale Room-Temperature Synthesis of High-Efficiency Lead-Free Perovskite Derivative (NH4)2SnCl6:Te Phosphor for Warm wLEDs. Chem. Eng. J. 2021, 420, 129740. [Google Scholar] [CrossRef]
  42. Giannozzi, P.; Baroni, S.; Bonini, N.; Calandra, M.; Car, R.; Cavazzoni, C.; Ceresoli, D.; Chiarotti, G.L.; Cococcioni, M.; Dabo, I. QUANTUM ESPRESSO: A modular and Ppen-Source Software Project for Quantum Simulations of Materials. J. Phys. Condens. Matter 2009, 21, 395502. [Google Scholar] [CrossRef] [PubMed]
Figure 1. (a) The crystal structure of Cs2ZnCl4. (b) The bond length of Zn-Cl and the bond angle of Cl-Zn-Cl in a [ZnCl4]2− tetrahedron before and after Zr4+ doping. (c) The XRD patterns of Cs2ZnCl4 with different Zr4+ feed ratios.
Figure 1. (a) The crystal structure of Cs2ZnCl4. (b) The bond length of Zn-Cl and the bond angle of Cl-Zn-Cl in a [ZnCl4]2− tetrahedron before and after Zr4+ doping. (c) The XRD patterns of Cs2ZnCl4 with different Zr4+ feed ratios.
Molecules 29 01651 g001
Figure 2. (a) Absorption spectra of Cs2ZnCl4 with different Zr4+ feed ratios. (b) Normalized PLE (with the monitoring wavelength at 457 nm) and PL (under the excitation wavelength at 260 nm) spectra of 20%Zr4+:Cs2ZnCl4. (c) PL lifetime of Cs2ZnCl4 with different Zr4+ feed ratios monitored at 260 nm. (d) Raman spectra of Cs2ZnCl4 with different Zr4+ feed ratios under the excitation of a 630 nm laser.
Figure 2. (a) Absorption spectra of Cs2ZnCl4 with different Zr4+ feed ratios. (b) Normalized PLE (with the monitoring wavelength at 457 nm) and PL (under the excitation wavelength at 260 nm) spectra of 20%Zr4+:Cs2ZnCl4. (c) PL lifetime of Cs2ZnCl4 with different Zr4+ feed ratios monitored at 260 nm. (d) Raman spectra of Cs2ZnCl4 with different Zr4+ feed ratios under the excitation of a 630 nm laser.
Molecules 29 01651 g002
Figure 3. (a) The map of the PL intensity of Zr4+:Cs2ZnCl4 plotted with temperature T and emission wavelength λ at the temperature range of 80–360 K. (b) Fitting results of exciton binding energy Eb. (c) Fitting results of Huang–Rhys factor S. (d) Fitting results of electron–optical-phonon coupling energy.
Figure 3. (a) The map of the PL intensity of Zr4+:Cs2ZnCl4 plotted with temperature T and emission wavelength λ at the temperature range of 80–360 K. (b) Fitting results of exciton binding energy Eb. (c) Fitting results of Huang–Rhys factor S. (d) Fitting results of electron–optical-phonon coupling energy.
Molecules 29 01651 g003
Figure 4. (a) Bandgap structure of undoped Cs2ZnCl4. (b) DOS of undoped Cs2ZnCl4. (c) Bandgap structure of Zr4+:Cs2ZnCl4. (d) DOS of Zr4+:Cs2ZnCl4.
Figure 4. (a) Bandgap structure of undoped Cs2ZnCl4. (b) DOS of undoped Cs2ZnCl4. (c) Bandgap structure of Zr4+:Cs2ZnCl4. (d) DOS of Zr4+:Cs2ZnCl4.
Molecules 29 01651 g004
Figure 5. Schematic diagram of the photophysical model of Zr4+:Cs2ZnCl4.
Figure 5. Schematic diagram of the photophysical model of Zr4+:Cs2ZnCl4.
Molecules 29 01651 g005
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Tian, Y.; Wei, Q.; Duan, L.; Peng, C. Boosting Blue Self-Trapped Exciton Emission in All-Inorganic Zero-Dimensional Metal Halide Cs2ZnCl4 via Zirconium (IV) Doping. Molecules 2024, 29, 1651. https://doi.org/10.3390/molecules29071651

AMA Style

Tian Y, Wei Q, Duan L, Peng C. Boosting Blue Self-Trapped Exciton Emission in All-Inorganic Zero-Dimensional Metal Halide Cs2ZnCl4 via Zirconium (IV) Doping. Molecules. 2024; 29(7):1651. https://doi.org/10.3390/molecules29071651

Chicago/Turabian Style

Tian, Ye, Qilin Wei, Lian Duan, and Chengyu Peng. 2024. "Boosting Blue Self-Trapped Exciton Emission in All-Inorganic Zero-Dimensional Metal Halide Cs2ZnCl4 via Zirconium (IV) Doping" Molecules 29, no. 7: 1651. https://doi.org/10.3390/molecules29071651

Article Metrics

Back to TopTop