REVIEW ARTICLE

Self-trapped exciton emission in inorganic copper(I) metal halides

  • Boyu ZHANG ,
  • Xian WU ,
  • Shuxing ZHOU ,
  • Guijie LIANG ,
  • Qingsong HU
Expand
  • Hubei Key Laboratory of Low Dimensional Optoelectronic Materials and Devices, Hubei University of Arts and Science, Xiangyang 441053, China

Received date: 21 Dec 2020

Accepted date: 02 Feb 2021

Published date: 15 Dec 2021

Copyright

2021 Higher Education Press

Abstract

The broad emission and high photoluminescence quantum yield of self-trapped exciton (STE) radiative recombination emitters make them an ideal solution for single-substrate, white, solid-state lighting sources. Unlike impurities and defects in semiconductors, the formation of STEs requires a lattice distortion, along with strong electron–phonon coupling, in low electron-dimensional materials. The photoluminescence of inorganic copper(I) metal halides with low electron-dimensionality has been found to be the result of STEs. These materials were of significant interest because of their lead-free, all-inorganic structures, and high luminous efficiencies. In this paper, we summarize the luminescence characteristics of zero- and one-dimensional inorganic copper(I) metal halides with STEs to provide an overview of future research opportunities.

Cite this article

Boyu ZHANG , Xian WU , Shuxing ZHOU , Guijie LIANG , Qingsong HU . Self-trapped exciton emission in inorganic copper(I) metal halides[J]. Frontiers of Optoelectronics, 2021 , 14(4) : 459 -472 . DOI: 10.1007/s12200-021-1133-4

1 Introduction

Halide perovskites have been recognized as a promising semiconductor material for next-generation optoelectronics, benefiting from high photoluminescence quantum yield (PLQY), tunable emission, and good charge mobility [1,2]. Photoelectric devices based on these materials, such as solar cells, light emitting devices, photodetectors, and lasers, have made unprecedented progress [2]. However, the long-term stability and toxicity of lead are major challenges to their commercialization [2]. To solve these problems, the strategy of replacing toxic lead with non-toxic ions (such as tin, germanium, antimony, and bismuth) has been extensively developed to maintain the advantageous properties of lead perovskites while improving the stability and reducing the toxicity [37]. However, some disadvantages still persist. Tin(II)-based perovskites have poor chemical stabilities, and Ge- or Bi-based metal halides have unsatisfactory photoelectric properties, which hamper their further optoelectronic applications [3,6,7].
Luo et al. reported a lead-free double perovskite, Cs2Ag(Na)In(Bi)Cl6, which is an important breakthrough [8]. This perovskite provides highly efficient and stable white light emission from self-trapped excitons (STEs). STEs are usually formed in semiconductors with low electron-dimensions and soft lattices [8]. Low electron-dimensionality is conducive to the carriers’ localization, and the soft lattice can elastically deform under photoexcitation, which leads to strong electron–phonon coupling [9]. STEs have a large exciton binding energy [9], which is one of their remarkable characteristics. Ordinary exciton luminescence has slightly lower energy than the bandgap, but as an unconventional exciton, STEs have much smaller emission energies than the bandgap [10], because the emission energy of STEs is equal to the difference between bandgap energy and the exciton binding energy, self-trapped energy, and lattice deformation energy, which leads to the large Stokes shift of the STEs emission energy [10].
STEs are found widely in halide crystals, condensed rare gases, and organic molecular crystals [11]. Since the discovery of STE emission in inorganic copper(I)-based metal halides (Cu(I)MHs), with Cu(I)-X (X= Cl, Br, I) polyhedra as the basic unit, further research of the STE emission from these materials has been rapidly developed [1217]. After Jun et al. first synthesized single crystals of Cs3Cu2I5 [12], the high PLQY (̴90%) and air stability of low electron-dimensional Cu(I)MHs immediately attracted great attention. Subsequently, reports regarding the regulation of the luminescence properties of such low electron-dimensional compounds by ion substitution have emerged. Through anion mixing, the emission of zero-dimensional (0D) Cs3Cu2X5 (X=Cl, Br, I) can be modulated to green light (in Cs3Cu2Cl5) from blue light (in Cs3Cu2I5), whereas the emission of the one-dimensional (1D) CsCu2X3 (X= Cl, Br, I) analog can be tuned from green to yellow [15,18,19]. However, the emission and PLQY of Cu(I)MHs show non-monotonic changes. The emission of 1D A2CuX3 (A= Rb, K; X= Br, Cl) is mainly concentrated in the ultraviolet region with a small tuning range [13,16,20]. The compound Cs5Cu3Cl6I2, another 1D Cu(I)MH, produces pure blue light [21]. By changing the anions and cations in the material, the bandgap, exciton binding energy, self-trapped energy, and lattice deformation energy are changed, resulting in the modulation of the emission color, Stokes shift, full width at half maximum (FWHM), luminescence lifetime, and other properties. In addition to color modulation, unusual luminescence phenomena including pressure-induced luminescence enhancement, polarization-dependent PL emission, and multi-photon absorption have also been observed [2224].
In this paper, we have summarized the current progress of inorganic Cu(I)MHs with STEs emission. The basic characteristics of STEs and their emission in these materials are described. The photophysical properties of STEs are explored from the perspective of the materials’ crystal structures, spectral properties, and electronic structures. We introduce the applications of these materials in optoelectronic devices. Consequently, we discuss the challenges faced in the future use of all-inorganic Cu(I)MHs.

2 STEs in 0D inorganic Cu(I)MH materials

Jun et al. reported the novel lead-free 0D luminescent material Cs3Cu2I5, which has a PLQY of nearly 90% for its single crystal and 60% for its thin film [12]. They attributed the emission to an excited-state structural reorganization mechanism caused by the Jahn-Teller distortion of tetrahedral Cu [12]. The inorganic and lead-free nature and efficient emission characteristics of Cs3Cu2I5 attracted great attention. As a typical 0D inorganic Cu(I)MH, Cs3Cu2X5 (X= Cl, Br, I) crystallizes in the orthorhombic structure with a Pnma space group. In this structure, a trigonal planar CuX3 shares an edge with a tetrahedral CuX4 unit to constitute a [Cu2X5]3 cluster. The [Cu2X5]3 dimers are completely separated by Cs+ cations, as shown in Fig. 1(a).
Fig.1 (a) Crystal structure of Cs3Cu2I5 (top), and Cs3Cu2Cl5 (middle). Two adjacent copper halide polyhedra (bottom) of Cs3Cu2I5 (left) and Cs3Cu2Br5 (right) along one of the crystal directions [25]. (b) X-ray diffraction (XRD) (left) and unit cell volumes (right) of different Cs3Cu2X5 (X= Cl, Br, I) compounds [18]. (c) PLE (left) and PL (right) spectra of different Cs3Cu2X5 [18]. (d) Images of the powders of different Cs3Cu2X5 (X= Cl, Br, I) compounds under ultraviolet irradiation [15,26]. (e) Integrated PL intensity vs. excitation power for Cs3Cu2Cl5 [26]. (f) Integrated PL intensity vs. excitation power for Cs3Cu2Br5−nIn(n = 0, 1.25, 2.50, 3.75) [15]. (g) Integrated PL intensity vs. excitation power for Cs3Cu2I5 [27]

Full size|PPT slide

Tab.1 PLE, PL, lifetime (t), PLQY, bandgap (Eg), FWHM, and Stokes shift (DS) of Cs3Cu2X5 (X= Cl, Br, I) reported in the literature
material PLE /nm PL/nm t/µs PLQY /% Eg /eV FWHM /nm DS/nm Ref.
Cs3Cu2Cl5 259 527 135.97 48.7 4.43 242 [17]
Cs3Cu2Cl5 ~283 516 78 ~235 [18]
Cs3Cu2Cl5 310 525 60 102 215 [25]
Cs3Cu2Cl5 320 515 112.4 91.3 3.60 91 195 [26]
Cs3Cu2(Cl0.75Br0.25)5 ~275 516 10 ~185 [18]
Cs3Cu2(Cl0.5Br0.5)5 ~272 470 3.6 ~195 [18]
Cs3Cu2(Cl0.25Br0.75)5 ~273 464 3.3 ~190 [18]
Cs3Cu2Br5 298 455 50.1 4.51 75 157 [15]
Cs3Cu2Br5 269 461 14.12 16.9 4.33 [17]
Cs3Cu2Br5 293 460 10 73 167 [25]
Cs3Cu2Br5 290 461 17.3 3.72 82 171 [26]
Cs3Cu2Br3.75I1.25 293 456 53.8 85 164 [15]
Cs3Cu2Br2.5I2.5 294 453 55.2 89 159 [15]
Cs3Cu2Br1.25I3.75 300 448 60.4 93 148 [15]
Cs3Cu2l5 290 445 91.2 175 155 [12]
Cs3Cu2l5 309 443 98.7 99 135 [15]
Cs3Cu2l5 262, 285 445 1.56 29.2 4.40 158 [17]
Cs3Cu2l5 310 440 62 80 130 [25]
Cs3Cu2l5 305 445 49.2 4.06 77 140 [26]
Cs3Cu2l5 286 444 1.12 4.1 79 [28]
Cs3Cu2l5 316 437 95 [29]
Sebastia-Luna et al. suggested that Cu(I)MHs can form a continuous solid solution by exchanging the halogen from Cl to Br and Br to I [18]. Their XRD patterns are fitted using a Pnma space group. The unit cell parameters increase with increasing anion radius (Cl<Br<I) (Fig. 1(b)) [18]. Correspondingly, the emission shifts to higher energy overall; however, the shift in maximum excitation and emission wavelengths are not monotonic when increasing the halide ionic radius (from Cl to I) as shown in Fig. 1(c), which is in agreement with the PL characteristics of Cs3Cu2Br5−nIn (0≤n≤5) reported in Refs. [15,18]. In general, the semiconductor bandgap decreases and the inter-band emission shifts to lower energy with increasing halogen anion radius. The emission shift trend here is different from that of inter-band radiation recombination, but analogous to that of CuX (X= Cl, Br, I) where the emission results from strong exciton-phonon coupling [15]. On the contrary, Cs3Cu2X5 shows very large Stokes shifts, as shown in Table 1. Typically, PL spectra with such large Stokes shifts are attributed to STEs as a result of strong exciton-phonon coupling [15]. To verify this inference, the excitation power dependences of the PL for Cs3Cu2X5 (X= Cl, Br, I) were measured [15,26,27]. As shown in Figs. 1(e)–1(g), the linear relationship between the emission intensity and the excitation power indicates the absence of permanent defects; hence, the emission likely originates from STEs [26]. There are some other effective experimental techniques to detect STE, such as characterizing the coupling strength of excitons and phonons, photoconversion spectroscopy [30]. Additionally, the broad excited-state transient absorption plateau across the probe region is a direct evidence of the formation of STEs [27]. Sebastia-Luna et al. suggested that the emission shift with the change of components may be caused by either halide segregation or phase segregation [18]. The XRD pattern of Cs3Cu2Cl5 is better fitted using the space group Cmcm. Compared to other Cs3Cu2X5 materials, Cs3Cu2Cl5 has a significantly red-shifted PL (516 nm) because of the different crystal structure. The dual excitation and emission properties of Cs3Cu2(Cl0.75Br0.25)5 suggest the coexistence of the Cmcm phase (as Cs3Cu2Cl5) and Pnma phase (as Cs3Cu2Br5) [18].
To obtain further information about the emission mechanism, the temperature-dependent steady and transient state PLs of Cs3Cu2Cl5 were measured over the range 300–20 K [26]. The results showed that the FWHM decreases gradually, the PL intensity increases, and the emission redshifts as the temperature decreases. The lack of free excitons emission with the change in temperature indicated that the self-trapped barrier is low and the photogenerated excitons are easily trapped [26]. The PL lifetimes of Cs3Cu2Cl5 increased significantly from 112.4 µs at 300 K to 877.7 µs at 20 K, while the mono-exponential decay was maintained. The mono-exponentiality of the lifetime indicates that there is no change in the radiative and nonradiative channels [26]. This temperature-dependent PL characterization is essential for the analysis of the relevant mechanism. It is commonly used to calculate the exciton binding energy and the S factor associated with the electro-phonon coupling strength [10]. The formation of STEs usually requires materials with large exciton binding energies and S values. A large S value means that there is strong electro-phonon coupling [10].
Fig.2 Projected density states for pristine structures of (a) Cs3Cu2Cl5, (b) Cs3Cu2Br5, and (c) Cs3Cu2I5 [26]. (d) Configuration coordinate diagram for the excitation, relaxation, and emission in Cs3Cu2Br5 [15]. (e) Configuration coordinate diagram for the formation of STEs in Cs3Cu2Cl5 [26]

Full size|PPT slide

Theoretical calculations found that the valence and conduction bands of Cs3Cu2X5 (X= Cl, Br, I) are mainly consist of Cu 3d and Cu 4s states (Figs. 2(a)–2(c)), indicating that they have narrow bandgaps [15,26]. Roccanova et al. suggests that the structural distortion breaks the two Cu-halogen bonds and moves the two Cu atoms closer to each other in the [Cu2X5]3− cluster, as shown in Fig. 2(d) [15]. The hole has Cu 3d character and is localized on the two Cu atoms in the [Cu2X5]3− cluster, whereas the electron is localized mostly in the bond-center position between the two Cu atoms [15]. The spatial distribution of the electron and hole of the exciton in Cs3Cu2I5 is similar to that in Cs3Cu2Br5 [15]. Lian et al. suggested that the photoexcitation induces a local structural distortion of Cs3Cu2X5 (X= Cl, Br, I), and the [Cu2X5]3− cluster is converted into a spindle-like species (two tetrahedrons sharing a face), decreasing the Cu-Cu distance as illustrated in Fig. 2(e) [26]. The distortion enhanced the local symmetry of the spindle-like [Cu2X5]3− and decreased the energy barrier for the formation of STEs, which explains the origin of the single emission from the STEs in Cs3Cu2X5 (X= Cl, Br, I) [26]. In this situation, the valence band maximum (VBM) shows stronger charge localization. However, as the anion goes down the halogen group, the increase of anionic p-component mixing with the Cu 3d orbitals leads to weakening of the VBM. This result partially explains the abnormal blue-shift in the emission of Cs3Cu2X5 (X= Cl, Br, I) when the anion changes from Cl to Br to I [26]. The calculations of Zhang et al. reveal that the electron and hole are separated spatially. This spatial separation impedes the recombination of the electron and hole and thus prolongs the PL lifetimes of Cs3Cu2X5 (X= Cl, Br, I) [15,26]. Because of the highly localized electron and hole, the emission of Cs3Cu2Cl5 is insensitive to defects, giving rise to the high PLQY and the mono-exponential PL decays.

3 STE in 1D inorganic Cu(I)MH materials

Three typical crystal structures of 1D Cu(I)MH are shown in Figs. 3(a)–3(f). Specifically, CsCu2X3 (X= Cl, Br, I) belongs to the orthorhombic Cmcm space group, where CuX4 tetrahedra share common edges, forming 1D chains separated by Cs+ ions [18]. Furthermore, A2CuX3 (A= Rb, K; X= Br, Cl) crystallizes in an orthorhombic structure with a Pnma space group, where CuX4 tetrahedra share a common corner, forming 1D chains separated by A+ ions along the b-axis [20]. The mixed halide compound Cs5Cu3Cl6I2 belongs to the orthorhombic Cmcm space group, where the alternately connected [CuCl2I2]2 unit and a single [CuCl2I2] unit form 1D zigzag chains of [Cu3Cl6I2]n5n, where only I ions bridge between these units as shown in Fig. 3(e). Similar to CsCu2X3, the 1D [Cu3Cl6I2]n5n chains were separated by Cs+ ions [21].
Fig.3 (a) Crystal structure of CsCu2I3 [31]. (b) CsCu2I3 structure viewed along the c-axis [31]. (c) Crystal structure of Rb2CuBr3 [20]. (d) Rb2CuBr3 structure viewed along the a-axis [20]. (e) Crystal structure of Cs5Cu3Cl6I2 [21]. (f) [Cu2I5]3− unit in Cs5Cu3Cl6I2 (Pnma) (left) and the [Cu2Cl5]3− unit in Cs5Cu3Cl6I2 (Cmcm) (right) [21]. (g) PL micrographs of CsCu2I3 crystal under pressures from 1 atm up to 16.0 GPa [22]. Schematic illustrations of the trapping and detrapping processes of the excitons in (h) phase I and (i) phase II. FE, free exciton state; GS, ground state; ST, self-trapped state; EST, self-trapping energy; EPL, photoluminescence energy; Ed, lattice deformation energy [22]

Full size|PPT slide

Tab.2 PLE, PL, t, PLQY, Eg, FWHM, and DS of CsCu2X3 (X= Cl, Br, I), A2CuX3 (A= Rb, K; X= Br, Cl), and Cs5Cu3Cl6I2 from the literatures
material PLE /nm PL/nm t/µs PLQY /% Eg/eV FWHM/nm DS/nm Ref.
CsCu2Cl3 319 527 0.0138 48.0 4.29 102 208 [19]
CsCu2Cl1.5Br1.5 340 587 0.0151 0.37 200 247 [19]
CsCu2Br3 319 533 0.018 18.3 3.94 106 214 [19]
CsCu2Br1.5I1.5 335 584 0.0266 0.38 128 249 [19]
CsCu2I3 347 568 0.0636 15.7 3.78 75 245 [14]
CsCu2I3 331 575 6 120 244 [25]
CsCu2I3 321 561 0.104 3.6 [28]
CsCu2I3 334 576 0.062 3.23 3.93 126 242 [19]
Rb2CuCl3 300 395 12.21 85 4.49 52 93 [13]
Rb2CuBr3 276 385 41.4 64 3.51 54 85 [13]
Rb2CuBr3 300 385 41.4 98.6 3.51 54 85 [20]
Rb2CuBr3 302 390 20 54 88 [25]
K2CuCl3 291 392 96.58 54 101 [16]
K2CuBr3 296 388 55 54 92 [16]
Cs5Cu3Cl6I2 462 40 95 95 191 [21]
In contrast to the blue-shift of the Cs3Cu2X5 (X= Cl, Br, I) emission with increasing anion radius, the emission of the single halide in CsCu2X3 (X= Cl, Br, I) shows a continuous red-shift with an increase in the Stokes shift [19]. However, the emission shifts of the mixed halides CsCu2Cl1.5Br1.5 and CsCu2Br1.5I1.5 do not follow a linear trend, and their maximum emission is red-shifted with lower PL efficiency compared to that of the single halide compounds (Table 2) [19]. Exceptionally, mixed Br–Cl compounds, which are rich in Cl, demonstrate minor phase segregation between CsCu2Br3 and CsCu2Cl3 [19]. For mixed halides, the emission behaviors are similar to those observed for MAPb(IBr)3 [19]. These facts could be explained by the presence of greater structural distortion in the mixed halide samples, which can affect the PL efficiency as well as the self-trapped depth within the bandgap of the material and result in a red-shift of the PL maximum [19]. Particularly, the PL lifetimes of CsCu2X3 (X= Cl, Br, I) increase from 13.8 ns for CsCu2Cl3 to 62 ns for CsCu2I3, following the trend opposite to the PL lifetimes of Cs3Cu2X5 (X= Cl, Br, I). The PL lifetimes of CsCu2X3 (X= Cl, Br, I) are also one to two orders of magnitude shorter than those of Cs3Cu2X5 (X= Cl, Br, I) [19,26]. The temperature-dependent PL shows that the emission of CsCu2I3 is continuously blue-shifted with increasing temperature [14]. A similar blue-shift was also observed in almost all CsCu2X3 (X= Cl, Br, I) compounds [19]. As mentioned above, such an abnormal blue-shift phenomenon is caused by the high lattice distortion and the electron–phonon renormalization.
To better understand STE emission in non-octahedral units, Li et al. investigated the pressure-induced PL enhancement in 1D Cu(I)MH CsCu2I3 [22]. The experimental results show that the slight structural distortion of the CuCl4 tetrahedra leads to bandgap broadening and STE enhancement under mild compression from 1 atm to 4.1 GPa. Such changes result in the increase of emission energy and intensity for CsCu2I3 in phase I. Meanwhile, the slight structural distortion results in a lower detrapping barrier, weaker electron-phonon coupling strength, and lower activation energy [22]. Therefore, the STE emission efficiencies of phase I are relatively low under this mild compression. The emission intensity is significantly enhanced with further continuous compression to 8.0 GPa. However, it is quenched with further compression beyond this limit, because the strong structural distortions lead to a phase transition and greatly increase the electron-phonon coupling, inducing a higher detrapping barrier and larger activation energy [22]. Thus, less STEs can be detrapped from the ST state to the free exciton (FE) state in phase II (Fig. 3(i)). This relatively increases the concentration of STEs, thus enhancing the STE emission [32,33]. The emission quenching phenomenon at high compression is usually attributed to the synergistic effect of deviatoric stress and structural amorphization [34].
Fig.4 (a) Partial charge density contours of both hole and electron wave functions for STE1 and STE2 in CsCu2Cl3 and STE3 in CsCu2I3, viewed from the two directions (along the b and a axes, respectively) perpendicular to the 1D chain [35]. (b) Femtosecond UV-vis transient absorption broadband spectra of a CsCu2I3 thin film [36]. Each curve in the figure represents the absorption curve of the CsCu2I3 film at a certain moment, and the vertical axis represents the change of absorption at a certain moment compared to the time zero. (c) Species-associated spectra were obtained from the kinetic modeling. Species: FE* and FE (hot and cooled free excitons, respectively), STE** and STE* (hot and partially cooled STEs, respectively). GS is the “ground-state” absorption (black solid line) [36]. (d) Semilogarithmic contour plots of experimental transient absorption spectra (right) and fitted transient absorption spectra (left), the latter with additional representations of the spectral evolution for each species [36]. (e) Energy level scheme summarizing the states, transitions, and central dynamic processes in CsCu2I3 thin films. Energetics of STE1, STE2, and STE3 are taken from Refs. [35,36]

Full size|PPT slide

The electronic structures of 1D Cu(I)MH materials are similar to the 0D Cu(I)MH mentioned above. The valence band (VB) and conduction band (CB) are predominantly composed of Cu 3d orbital and mixed Cu 4s orbitals. The M+ (Cs+, Rb+, K+, etc.) ions are far away from the VB and CB, so their contribution to the energy band can be ignored [13,15,16,20,26]. In addition, the electronic structures of Cu(I)MHs significantly change upon photoexcitation, which is strongly correlated with their optical properties [26]. Du also found, using first principles calculations, that the emission energy of CsCu2Cl3, CsCu2Br3, and CsCu2I3 STEs decreased gradually as the anion goes down the halogen group [35]. The red-shift trend experimentally observed for the emission spectra of the three materials is not caused by the electronegativity of anions, but rather from the different STE properties [35]. Du’s calculation showed that each CsCu2X3 (X= Cl, Br, I) has three different STE states. Figure 4(a) shows the partial charge density contours of both the hole and electron wave functions for STE1 and STE2 in CsCu2Cl3 and STE3 in CsCu2I3. The calculated emission of CsCu2Br3 and CsCu2I3 originates from the irradiative transfer of the lowest energy level of the STE, whereas CsCu2Cl3 emits from the STE’s metastable energy level [35]. These results are in excellent agreement with the experimental results. Based on Du’s calculations, Kentsch et al. suggest that the exciton self-trapped process in CsCu2I3 films was directly observed by femtosecond transient absorption [36]. The edge absorption bimodal structure is derived from the 130 meV spin orbital splitting of the Cu d electrons. The formation of the lowest level of the STE state results in the disappearance of band edge absorption. The experiment showed that the time constant of the self-trapped process of the thermal relaxation free exciton is 12 ps and the energy barrier is not lower than 60 meV [36]. Each process is obtained by species-associated spectra synthesis. It is probably the most powerful experimental description of STE at present.
The emission from CsCu2X3 (X= Cl, Br, I) has a large FWHM and Stokes shift, similar to the above-mentioned characteristics of the STEs in Cs3Cu2X5 (X= Cl, Br, I), but it exhibits the shorter PL lifetime. It is unusual that although a STE mechanism is observed, the FWHM and Stokes shift of the A2CuX3 (A= Rb, K; X= Cl, Br) emission are much smaller than those of CsCu2X3 (X= Cl, Br, I). In addition, Rb2CuCl3 showed an upconversion emission phenomenon [13]. In this process, photons with lower energy than the bandgap interact with the lattice, causing absorption of phonons in the lattice and subsequent emission of photons with higher energy than before. This mechanism leads to the extraction of heat energy from the lattice to emit the higher energy photon, which cools down the material. The advantages of high PLQY and low trapped state density in these Cu(I)MHs are important parameters for optical cooling [13]. Among all the reported inorganic 1D Cu(I)MHs, Cs5Cu3Cl6I2 was the only one to emit pure blue emission, with a near-unity quantum yield of ≈95%. The flat valence bands promote the formation of localized holes, leading to STE formation. The photo-deformable nature of the copper halogen polyhedra facilitates highly efficient STEs, leading to efficient radiative recombination [21].

4 Applications of inorganic Cu(I)MH STE materials

Fig.5 (a) Photo of a white LED driven under a bias voltage of 6 V. The inset shows photographs of Cs3Cu2I5, a Cs3Cu2I5 + CsCu2I3 mixture, and CsCu2I3 (from top to bottom) under UV light irradiation at 305 nm [37]. (b) PL spectrum of the Cs3Cu2I5 + CsCu2I3 mixture (with weight ratio of 1:16 or molar ratio of 1:8). Inset shows a photograph of the mixture under 254 nm light irradiation [25]. (c) Schematic structure of the CsCu2I3 based LEDs [38]. (d) Voltage-current density-luminance curves of the yellow emission LEDs [38]. (e) Schematic structure of the Cs3Cu2I5 NCs based LEDs [39]. (f) Voltage-current density-luminance curves of the blue emission device [39]. (g) Schematic structure of the CsCu2I3@Cs3Cu2I5 based LEDs [40]. (h) Voltage-current density-luminance curves of the white emission device [40]

Full size|PPT slide

Based on the advantages mentioned above, inorganic Cu(I)MH STE materials are very suitable for applications in lighting and displays as a phosphor. Cs3Cu2I5 and CsCu2I3 (Fig. 5) can coexist in the synthesis process; this gives the opportunity to obtain tunable white light emission using a simple mechanochemical method to quickly synthesize mixed-phase materials with different proportions [37]. Fang et al. used this method to obtain white luminescent samples by mixing CsCu2I3 and Cs3Cu2I5 with different ratios. When the mixture had a mass ratio of 8:3, the sample emitted with a Commission Internationale de L’Eclairage (CIE) chromaticity coordinate of (0.36, 0.36) [37]. In other reports, pure CsCu2I3 and Cs3Cu2I5 materials (powders and nanocrystals) were synthesized first and then mixed in suitable proportions to produce white light emitters [25,28]. Furthermore, CsCu2I3 single crystal (SC) has great potential in energy-saving white lighting because of its bimodal emission [14,41]. However, ensuring the stability of the light efficiency and color at different temperatures is a key problem to be solved in the lighting and display applications of such materials.
Inorganic Cu(I)MH STE materials can also be directly used as an electroluminescent layer in LED devices. Ma et al. fabricated CsCu2I3-based LEDs with an ITO/PEDOT:PSS/poly-TPD/CsCu2I3/TPBi/LiF/Al structure [38]. The best device had a turn-on voltage of about 5.0 V, a maximum luminance (Lmax) value of 47.5 cd/m2 at an applied bias voltage of 9.2 V, and maximum external quantum efficiency (EQE) of 0.17% [38]. Then they constructed a high color-rendering index (CRI= 91.6) and stable white light-emitting diode (WLED) with an ITO/PEDOT:PSS/poly-TPD/PVK/CsCu2I3@Cs3Cu2I5/TPBi/LiF/Al structure [40]. Liu et al. also reported a WLED with an ITO/PEDOT:PSS/V-NPB/CsCu2I3@Cs3Cu2I5/SPPO13/LiF/Al structure. This WLED had a CIE chromaticity coordinate of (0.327, 0.348) and a CRI of up to 94 [29]. Wang et al. reported deep-blue LEDs using Cs3Cu2I5 nanocrystals as the electroluminescent layer [39]. These blue LEDs had a glass/ITO/p-NiO/Cs3Cu2I5/TPBi/LiF/A structure, a turn-on voltage of about 4.5 V, a Lmax value of 262.6 cd/m2 at 7.5 V, and an EQE of ~1.12% at an applied bias voltage of 7.5 V [39]. These results indicated that copper(I) halides are promising as an environment-friendly and stable emitter for LEDs and are compatible with the practical applications of the devices.
Fig.6 (a) Schematic illustration of the Cs3Cu2I5/GaN heterojunction device [42]. (b) IV curves of the photodetector tested in the dark and under different light irradiation intensities (320 nm) [42]. (c) Responsivity and specific detectivity of the photodetector versus light intensity [42]. (d) Rising and falling edges for estimating the rise time (tr) and fall time (tf) of the photodetector [42]. (e) Schematic illustration of the polarization-sensitive photodetector based on 1D CsCu2I3 nanowires (NWs) [23]. (f) Anisotropic photocurrent response under 325 nm light excitation described via a 2D color map (photocurrent is denoted by the color bar, with voltage as the x-axis and polarization angle as the y-axis) [23]. (g) Photocurrent response of a 1D CsCu2I3 NW device under incident light with different polarization angles [23]. (h) and (i) Optical microscopy images of CsCu2I3 single crystals with rod-shaped morphology [44]. Comparison of photoelectric properties between (110) and (010) crystal planes of CsCu2I3 single crystals: (j) I–t curves and (k) I–V curves in dark and under 350 nm illumination at a 3 V bias [44]

Full size|PPT slide

Photodetection is another important application of these high-efficiency inorganic Cu(I)MH STE materials. Li et al. reported Cs3Cu2I5-based ultraviolet photodetectors with high spectral selectivity [42]. The Cs3Cu2I5/GaN heterojunction device had a narrow spectral response “window” of 300–370 nm, and the response range can be adjusted by changing the thickness of Cs3Cu2I5 film and bias voltage [42]. At zero bias, the responsivity, detection rate, maximum current on/off ratio, and response speeds were 0.28 A/W, 1.4 × 1012 Jones, 1.2 × 105, and 95/130 µs, respectively [42]. These values are comparable to those of the numerous previously reported lead halide photodetectors [42]. In addition, Zhang et al. reported a Cs3Cu2I5-based deep ultraviolet photodetector with a responsivity of 64.9 mA/W, a specific detectivity of 6.9 × 1011 Jones, a turn-on ratio of 127, and response speeds of 26.2/49.9 ms for the rise/fall time [43]. Li et al. also constructed a polarization-sensitive and flexible ultraviolet photodetector using 1D CsCu2I3 nanowires, which had a high photocurrent anisotropy ratio up to 3.16 because of the electrical and optical anisotropy of the asymmetric structure and the external morphological anisotropy [23]. The device had an on/off ratio of 2.6 × 103, a response speed of 6.94/214 µs, a photoresponsivity of 32.3 A/W, and a specific detectivity of 1.89×1012 Jones [23]. Similarly, Yang et al. fabricated a deep ultraviolet photodetector based on a CsCu2I3 film [31]. The on/off ratio, responsivity, specific detectivity, and EQE of this device were 22, 22.1 mA/W, 1.2 × 1011 Jones, and 10.3%, respectively, under 265 nm illumination with a light density of 0.305 mW/cm2 [31]. Furthermore, Fang et al. observed the facet-dependent photoresponse of a CsCu2I3 single crystal, whose morphology consists of 010, 110, and 021 crystal planes (Fig. 6) [44]. The lower dark current of the 010 crystal plane results in the on-off ratio being higher than that of 110 crystal plane [44]. This work gives new insights into the 1D electronic structure associated with high anisotropy.
Fig.7 (a) Schematic of the prototype projection system for X-ray imaging, and photographs of an X-ray image of a universal board and a ball-point pen [47]. (b) Illustration of Cs3Cu2I5 precursor solution as a fluorescent ink [48]. (c) Schematic illustration of the setup for recording deep UV images and the image-sensing profile of HFUT under 265, 365, and 405 nm light excitation [43]

Full size|PPT slide

Other possible applications of inorganic Cu(I)MH STE materials have also been explored, such as X-ray scintillators, fluorescent inks, image sensors, and memristors and neuromorphic computing applications [20,43,4549]. The excellent luminescence performance and lack of self-absorption in RbCu2Br3, RbCu2Cl3, K2CuBr3, and Cs3Cu2I5 led to a good scintillation response to X-ray signals, with a high light yield of 91056, 16600, 23806, and 79279 photons per meV, respectively [20,4547]. Zhang et al. reported a fluorescent ink developed from a Cs3Cu2I5/polyvinylidene fluoride precursor solution. This work showed potential applications of Cs3Cu2I5 for anti-counterfeiting and encryption fields (Fig. 7) [48]. Zhang et al. explored the image-sensing capability of Cs3Cu2I5 crystalline films, where purple and light purple images can be readily formed under 265 and 365 nm illuminations, respectively [43]. Furthermore, Zeng et al. investigated the application of Cs3Cu2I5 films in the field of memristors and neuromorphic computing [49]. Ag/PMMA/Cs3Cu2I5/ITO structure memristors exhibited a bipolar resistive switching operating voltage lower than ±1 V, a large on/off ratio (up to 102), stable endurance over 100 cycles, and a retention time longer than 104 s [49].

5 Conclusions and perspective

We have summarized the recent progress made in the field of STE emission from inorganic Cu(I)MHs. Novel 0D and 1D inorganic Cu(I)MHs STE luminescent materials, such as Cs3Cu2X5 (X= Cl, Br, I), CsCu2X3 (X= Cl, Br, I), A2CuX3 (A= Rb, K; X= Cl, Br), and Cs5Cu3Cl6I2, have been developed and researched. The basic features of STE luminescence have been summarized, methods for identifying STE luminescence have been proposed, and the application of these materials has been introduced. Evidently, the STE luminescence can be regulated by substituting cations and anions. In the process of ion modulation, the emission and light efficiency usually change non-monotonically, but the essential reason for this is currently unclear. Among the 1D materials, only CsCu2X3 has a PL lifetime of the order of nanoseconds, whereas other materials have a PL lifetimes of the order of microseconds. In addition, A2CuX3 (A= Rb, K; X= Cl, Br) have incomprehensible small emission FWHMs and Stokes shifts. These issues have not been thoroughly studied. Meanwhile, the formation process of STEs in this kind of material is rarely studied, and the direct experimental observation has not been realized. Therefore, the modulation mechanism has not been completely understood. A profound understanding of STE formation processes and modulation mechanisms is conducive to the accurate judgment and regulation of material properties and the development of similar materials with desired target properties. These studies will be essential for the performance optimization and extend the application of inorganic Cu(I)MH STEs and similar materials.

Acknowledgements

Q. Hu acknowledges the support from the National Key Research and Development Plan of China (No. 2019YFE0107200), the National Natural Science Foundation of China (Grant No. 11705277), the Natural Science Foundation of Hubei Province (No. 2020CFB700), the Doctoral Research Foundation Project of Hubei University of Arts and Science (No. kyqdf2020023), and Innovation Research Team Project of Hubei University of Arts and Science (No. 2020kypytd001).
1
Guo Z, Li J, Pan R, Cheng J, Chen R, He T. All-inorganic copper(I)-based ternary metal halides: promising materials toward optoelectronics. Nanoscale, 2020, 12(29): 15560–15576

DOI PMID

2
Tang J, Li D. Halide perovskites: from materials to optoelectronic devices. Frontiers of Optoelectronics, 2020, 13(3): 191–192

DOI

3
Wang L, Chen P, Kuttipillai P S, King I, Staples R, Sun K, Lunt R R. Epitaxial stabilization of tetragonal cesium tin iodide. ACS Applied Materials & Interfaces, 2019, 11(35): 32076–32083

DOI PMID

4
Chen J, Luo Z, Fu Y, Wang X, Czech K J, Shen S, Guo L, Wright J C, Pan A, Jin S. Tin(IV)-tolerant vapor-phase growth and photophysical properties of aligned cesium tin halide perovskite (CsSnX3; X= Br, I) nanowires. ACS Energy Letters, 2019, 4(5): 1045–1052

DOI

5
Wong A B, Bekenstein Y, Kang J, Kley C S, Kim D, Gibson N A, Zhang D, Yu Y, Leone S R, Wang L W, Alivisatos A P, Yang P. Strongly quantum confined colloidal cesium tin iodide perovskite nanoplates: lessons for reducing defect density and improving stability. Nano Letters, 2018, 18(3): 2060–2066

DOI PMID

6
Chen L J. Synthesis and optical properties of lead-free cesium germanium halide perovskite quantum rods. RSC Advances, 2018, 8(33): 18396–18399

DOI

7
Leng M, Yang Y, Zeng K, Chen Z, Tan Z, Li S, Li J, Xu B, Li D, Hautzinger M P, Fu Y, Zhai T, Xu L, Niu G, Jin S, Tang J. All-inorganic bismuth-based perovskite quantum dots with bright blue photoluminescence and excellent stability. Advanced Functional Materials, 2018, 28(1): 1704446

DOI

8
Luo J, Wang X, Li S, Liu J, Guo Y, Niu G, Yao L, Fu Y, Gao L, Dong Q, Zhao C, Leng M, Ma F, Liang W, Wang L, Jin S, Han J, Zhang L, Etheridge J, Wang J, Yan Y, Sargent E H, Tang J. Efficient and stable emission of warm-white light from lead-free halide double perovskites. Nature, 2018, 563(7732): 541–545

DOI PMID

9
Li J, Wang H, Li D. Self-trapped excitons in two-dimensional perovskites. Frontiers of Optoelectronics, 2020, 13(3): 225–234

DOI

10
Li S, Luo J, Liu J, Tang J. Self-trapped excitons in all-inorganic halide perovskites: fundamentals, status, and potential applications. Journal of Physical Chemistry Letters, 2019, 10(8): 1999–2007

DOI PMID

11
Smith M D, Karunadasa H I. White-light emission from layered halide perovskites. Accounts of Chemical Research, 2018, 51(3): 619–627

DOI PMID

12
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. Advanced Materials, 2018, 30(43): e1804547

DOI PMID

13
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. Advanced Optical Materials, 2020, 8(2): 1901338

DOI

14
Lin R, Guo Q, Zhu Q, Zhu Y, Zheng W, Huang F. All-inorganic CsCu2I3 single crystal with high-PLQY (≈15.7%) intrinsic white-light emission via strongly localized 1D excitonic recombination. Advanced Materials, 2019, 31(46): e1905079

DOI PMID

15
Roccanova R, Yangui A, Nhalil H, Shi H, Du M H, Saparov B. Near-unity photoluminescence quantum yield in blue-emitting Cs3Cu2Br5−xIx (0≤x≤5). ACS Applied Electronic Materials, 2019, 1(3): 269–274

DOI

16
Creason T D, McWhorter T M, Bell Z, Du M H, Saparov B. K2CuX3 (X= Cl, Br): all-inorganic lead-free blue emitters with near-unity photoluminescence quantum yield. Chemistry of Materials, 2020, 32(14): 6197–6205

DOI

17
Luo Z, Li Q, Zhang L, Wu X, Tan L, Zou C, Liu Y, Quan Z. 0D Cs3Cu2X5 (X= I, Br, and Cl) nanocrystals: colloidal syntheses and optical properties. Small, 2020, 16(3): e1905226

DOI PMID

18
Sebastia-Luna P, Navarro-Alapont J, Sessolo M, Palazon F, Bolink H J. Solvent-free synthesis and thin-film deposition of cesium copper halides with bright blue photoluminescence. Chemistry of Materials, 2019, 31(24): 10205–10210

DOI

19
Roccanova R, Yangui A, Seo G, Creason T D, Wu Y, Kim D Y, Du M H, Saparov B. Bright luminescence from nontoxic CsCu2X3 (X= Cl, Br, I). ACS Materials Letters, 2019, 1(4): 459–465

20
Yang B, Yin L, Niu G, Yuan J H, Xue K H, Tan Z, Miao X S, Niu M, Du X, Song H, Lifshitz E, Tang J. Lead-free halide Rb2CuBr3 as sensitive X-ray scintillator. Advanced Materials, 2019, 31(44): e1904711

DOI PMID

21
Li J, Inoshita T, Ying T, Ooishi A, Kim J, Hosono H. A highly efficient and stable blue-emitting Cs5Cu3Cl6I2 with a 1D chain structure. Advanced Materials, 2020, 32(37): e2002945

DOI PMID

22
Li Q, Chen Z, Yang B, Tan L, Xu B, Han J, Zhao Y, Tang J, Quan Z. Pressure-induced remarkable enhancement of self-trapped exciton emission in one-dimensional CsCu2I3 with tetrahedral units. Journal of the American Chemical Society, 2020, 142(4): 1786–1791

DOI PMID

23
Li Y, Shi Z, Wang L, Chen Y, Liang W, Wu D, Li X, Zhang Y, Shan C, Fang X. Solution-processed one-dimensional CsCu2I3 nanowires for polarization-sensitive and flexible ultraviolet photodetectors. Materials Horizons, 2020, 7(6): 1613–1622

DOI

24
Guo Z, Li J, Gao Y, Cheng J, Zhang W, Pan R, Chen R, He T. Multiphoton absorption in low-dimensional cesium copper iodide single crystals. Journal of Materials Chemistry C, Materials for Optical and Electronic Devices, 2020, 8(47): 16923–16929

DOI

25
Grandhi G K, Viswanath N S M, Cho H B, Han J H, Kim S M, Choi S, Im W B. Mechanochemistry as a green route: synthesis, thermal stability, and postsynthetic reversible phase transformation of highly-luminescent cesium copper halides. Journal of Physical Chemistry Letters, 2020, 11(18): 7723–7729

DOI PMID

26
Lian L, Zheng M, Zhang P, Zheng Z, Du K, Lei W, Gao J, Niu G, Zhang D, Zhai T, Jin S, Tang J, Zhang X, Zhang J. Photophysics in Cs3Cu2X5 (X= Cl, Br, or I): highly luminescent self-trapped excitons from local structure symmetrization. Chemistry of Materials, 2020, 32(8): 3462–3468

DOI

27
Cheng P, Sun L, Feng L, Yang S, Yang Y, Zheng D, Zhao Y, Sang Y, Zhang R, Wei D, Deng W, Han K. Colloidal synthesis and optical properties of all-inorganic low-dimensional cesium copper halide nanocrystals. Angewandte Chemie International Edition, 2019, 58(45): 16087–16091

DOI PMID

28
Vashishtha P, Nutan G V, E. Griffith B, Fang Y, Giovanni D, Jagadeeswararao M, Sum T C, Mathews N, Mhaisalkar S G, Hanna J V, White T. Cesium copper iodide tailored nanoplates and nanorods for blue, yellow, and white emission. Chemistry of Materials, 2019, 31(21): 9003–9011

DOI

29
Liu S, Yue Y, Zhang X, Wang C, Yang G, Zhu D. A controllable and reversible phase transformation between all-inorganic perovskites for white light emitting diodes. Journal of Materials Chemistry C, Materials for Optical and Electronic Devices, 2020, 8(25): 8374–8379

DOI

30
Williams R T, Song K S. The self-trapped exciton. Journal of Physics and Chemistry of Solids, 1990, 51(7): 679–716

DOI

31
Yang J, Kang W, Liu Z, Pi M, Luo L B, Li C, Lin H, Luo Z, Du J, Zhou M, Tang X. High-performance deep ultraviolet photodetector based on a one-dimensional lead-free halide perovskite CsCu2I3 film with high stability. Journal of Physical Chemistry Letters, 2020, 11(16): 6880–6886

DOI PMID

32
Ma Z, Liu Z, Lu S, Wang L, Feng X, Yang D, Wang K, Xiao G, Zhang L, Redfern S A T, Zou B. Pressure-induced emission of cesium lead halide perovskite nanocrystals. Nature Communications, 2018, 9(1): 4506

DOI PMID

33
Shi Y, Ma Z, Zhao D, Chen Y, Cao Y, Wang K, Xiao G, Zou B. Pressure-induced emission (PIE) of one-dimensional organic tin bromide perovskites. Journal of the American Chemical Society, 2019, 141(16): 6504–6508

DOI PMID

34
Wang Y, Lü X, Yang W, Wen T, Yang L, Ren X, Wang L, Lin Z, Zhao Y. Pressure-induced phase transformation, reversible amorphization, and anomalous visible light response in organolead bromide perovskite. Journal of the American Chemical Society, 2015, 137(34): 11144–11149

DOI PMID

35
Du M H. Emission trend of multiple self-trapped excitons in luminescent 1D copper halides. ACS Energy Letters, 2020, 5(2): 464–469

DOI

36
Kentsch R, Morgenroth M, Scholz M, Xu K, Schmedt Auf der Günne J, Lenzer T, Oum K. Direct observation of the exciton self-trapped process in CsCu2I3 thin films. Journal of Physical Chemistry Letters, 2020, 11(11): 4286–4291

DOI PMID

37
Fang S, Wang Y, Li H, Fang F, Jiang K, Liu Z, Li H, Shi Y. Rapid synthesis and mechanochemical reactions of cesium copper halides for convenient chromaticity tuning and efficient white light emission. Journal of Materials Chemistry. C, Materials for Optical and Electronic Devices, 2020, 8(14): 4895–4901

DOI

38
Ma Z, Shi Z, Qin C, Cui M, Yang D, Wang X, Wang L, Ji X, Chen X, Sun J, Wu D, Zhang Y, Li X J, Zhang L, Shan C. Stable yellow light-emitting devices based on ternary copper halides with broadband emissive self-trapped excitons. ACS Nano, 2020, 14(4): 4475–4486

DOI PMID

39
Wang L, Shi Z, Ma Z, Yang D, Zhang F, Ji X, Wang M, Chen X, Na G, Chen S, Wu D, Zhang Y, Li X, Zhang L, Shan C. Colloidal synthesis of ternary copper halide nanocrystals for high-efficiency deep-blue light-emitting diodes with a half-lifetime above 100 h. Nano Letters, 2020, 20(5): 3568–3576

DOI PMID

40
Ma Z, Shi Z, Yang D, Li Y, Zhang F, Wang L, Chen X, Wu D, Tian Y, Zhang Y, Zhang L, Li X, Shan C. High color-rendering index and stable white light-emitting diodes by assembling two broadband emissive self-trapped excitons. Advanced Materials, 2021, 33(2): e2001367

DOI PMID

41
Lin R, Zhu Q, Guo Q, Zhu Y, Zheng W, Huang F. Dual self-trapped exciton emission with ultrahigh photoluminescence quantum yield in CsCu2I3 and Cs3Cu2I5 perovskite single crystals. Journal of Physical Chemistry C, 2020, 124(37): 20469–20476

DOI

42
Li Y, Shi Z, Liang W, Wang L, Li S, Zhang F, Ma Z, Wang Y, Tian Y, Wu D, Li X, Zhang Y, Shan C, Fang X. Highly stable and spectrum-selective ultraviolet photodetectors based on lead-free copper-based perovskites. Materials Horizons, 2020, 7(2): 530–540

DOI

43
Zhang Z X, Li C, Lu Y, Tong X W, Liang F X, Zhao X Y, Wu D, Xie C, Luo L B. Sensitive deep ultraviolet photodetector and image sensor composed of inorganic lead-free Cs3Cu2I5 perovskite with wide bandgap. Journal of Physical Chemistry Letters, 2019, 10(18): 5343–5350

DOI PMID

44
Li Z, Li Z, Shi Z, Fang X. Facet-dependent, fast response, and broadband photodetector based on highly stable all-inorganic CsCu2I3 single crystal with 1d electronic structure. Advanced Functional Materials, 2020, 30(28): 2002634

DOI

45
Zhao X, Niu G, Zhu J, Yang B, Yuan J H, Li S, Gao W, Hu Q, Yin L, Xue K H, Lifshitz E, Miao X, Tang J. All-inorganic copper halide as a stable and self-absorption-free X-ray scintillator. Journal of Physical Chemistry Letters, 2020, 11(5): 1873–1880

DOI PMID

46
Gao W, Niu G, Yin L, Yang B, Yuan J H, Zhang D, Xue K H, Miao X, Hu Q, Du X, Tang J. One-dimensional all-inorganic K2CuBr3 with violet emission as efficient X-ray scintillators. ACS Applied Electronic Materials, 2020, 2(7): 2242–2249

DOI

47
Lian L, Zheng M, Zhang W, Yin L, Du X, Zhang P, Zhang X, Gao J, Zhang D, Gao L, Niu G, Song H, Chen R, Lan X, Tang J, Zhang J. Efficient and reabsorption-free radioluminescence in Cs3Cu2I5 nanocrystals with self-trapped excitons. Advanced Science, 2020, 7(11): 2000195

DOI PMID

48
Zhang F, Zhao Z, Chen B, Zheng H, Huang L, Liu Y, Wang Y, Rogach A L. Strongly emissive lead-free 0D Cs3Cu2I5 perovskites synthesized by a room temperature solvent evaporation crystallization for down-conversion light-emitting devices and fluorescent inks. Advanced Optical Materials, 2020, 8(8): 1901723

DOI

49
Zeng F, Guo Y, Hu W, Tan Y, Zhang X, Feng J, Tang X. Opportunity of the lead-free all-inorganic Cs3Cu2I5 perovskite film for memristor and neuromorphic computing applications. ACS Applied Materials & Interfaces, 2020, 12(20): 23094–23101

DOI PMID

Outlines

/