Open Access Article
Ulises Eli
Cercas
a,
Diego
Solis-Ibarra
b,
Fernando
Cortés-Guzman
c,
Iván Enrique
Martínez-Merlin
a,
Vojtech
Jancik
de,
Juan Carlos
Alonso-Huitrón
b,
Disnel
Ferrera-Carracedo
b,
Ciro
Falcony-Guajardo
f,
María
de Jesús Rosales-Hoz
g and
Jesús Uriel
Balderas-Aguilar
*b
aTecnológico Nacional de México, Instituto Tecnológico de Tlalnepantla, DEPI, Av. Instituto Tecnológico S/N, Col. La Comunidad, Tlalnepantla de Baz 54070, Mexico
bInstituto de Investigaciones en Materiales, Universidad Nacional Autónoma de México, Ciudad Universitaria, A.P. 70-360, Coyoacán, Mexico City, Mexico. E-mail: jubalderas@iim.unam.mx
cFacultad de Química, Universidad Nacional Autónoma de México, Ciudad Universitaria, CDMX 04510, Mexico
dInstituto de Química, Universidad Nacional Autónoma de México, Ciudad Universitaria, Ciudad de México 04510, Mexico
eCentro Conjunto de Investigación en Química Sustentable UAEM-UNAM, Carretera Toluca-Atlacomulco km. 14.5, Toluca, Estado de México 50200, Mexico
fCentro de Investigación y de Estudios Avanzados del IPN (CINVESTAV), Departamento de Física, Av. Instituto Politécnico Nacional 2508, San Pedro Zacatenco, Gustavo A. Madero, CDMX, Mexico
gCentro de Investigación y de Estudios Avanzados del IPN (CINVESTAV), Departamento de Química, Av. Instituto Politécnico Nacional 2508, San Pedro Zacatenco, Gustavo A. Madero, CDMX, Mexico
First published on 2nd October 2025
Zero-dimensional copper halide phosphors have emerged as efficient, less toxic alternatives to lead-based systems. Here, we report a bromide–iodide alloyed cluster, TPA2[Cu4Br2I4], stabilized by optoelectronically inert tetrapropylammonium (TPA+), featuring strong Cu–Cu interactions, visible-light excitation (λex = 444 nm), green-yellow emission (λem = 512 nm), and a high photoluminescence quantum yield (∼95%). Temperature- and power-dependent photoluminescence studies reveal a dual emission mechanism arising from excitonic recombination in cis and trans isomers. Time-dependent DFT calculations elucidate the photoinduced electronic redistribution and structural relaxation underlying this behavior. The material is readily synthesized via mechanochemistry and processed into thin films by spin coating or aerosol-assisted chemical vapor deposition (AACVD). Its integration into polymer composites demonstrates its practical utility for lighting applications. Finally, the successful extension of the alloying strategy to other cations such as MTPP+ highlights its versatility for designing tunable copper cluster halide phosphors aimed at next-generation optoelectronic and lighting technologies.
To mitigate the toxicity of lead, copper(I) halides have emerged as promising alternatives. Their luminescence is strongly influenced by coordination geometry (tetrahedral, trigonal planar, or linear), distortions, and Cu⋯Cu interactions. Most reported Cu(I) halides exhibit self-trapped exciton (STE)-driven emission, achieving high PLQY but requiring high-energy excitation (>3.5 eV), which exceeds the operating range of cost-effective blue-light LEDs and lasers commonly used in modern devices.3,4
Recently, 0D luminescent copper cluster halides containing [Cu4I6]2− subunits, stabilized by bulky organic cations, have demonstrated visible-light excitation.5–8 Metallophilic interactions within Cu4 clusters, arising from Cu⋯Cu distances shorter than their van der Waals radii, reduce the HOMO–LUMO gap, enabling excitation in the visible range. Furthermore, organic spacers – unlike inorganic cations – offer tunability in optical properties and luminescence through band alignment and weak interactions, making PLQY highly dependent on spacer selection. For example, highly conjugated methyltriphenylphosphonium spacers facilitate metal-to-ligand charge transfer (MLCT) and halide-to-ligand charge transfer (XLCT) mechanisms that contribute to luminescence, but at the cost of reduced PLQY.9 In contrast, non-conjugated tetraalkylammonium cations have been used to stabilize [Cu4Br6]2− clusters, yielding ultrabroad red-peaking luminescence with near-unity PLQY; however, in this system the excitation peaks at λ < 400 nm,10,11 significantly limiting its activation under visible-light sources. During the preparation of this manuscript, an independent study reported TPA2[Cu4Br2I4] as a down-converting and UV-protective layer in perovskite solar cells.12 Initially poor film morphology obtained from antisolvent-assisted spin-coating (DMF solutions) was greatly improved by vacuum-assisted dry spin-coating and post-annealing. Optimized devices retained 90% and 80% of their initial power conversion efficiencies after UV exposures of 66.07 and 210.7 kWh m−2, respectively, with a slight drop in efficiency (23.09% to 22.24%) due to reduced short-circuit current density. This highlights the broad utility and potential of the mixed-halide approach.
Here, we significantly advance the understanding and practical utilization of bromide–iodide alloyed copper-cluster halides. Single crystals synthesized from an optimized CuI
:
TPA-Br (4
:
2) precursor ratio yield [Cu4Br2I4]2− luminescent clusters with an exceptionally broad excitation band centered at 444 nm and bright green-yellow emission at 513 nm, accompanied by a remarkably high photoluminescence quantum yield (∼95%). Structural analysis via single-crystal X-ray diffraction, complemented by electronic structure calculations, provides detailed insight into the excited-state dynamics responsible for the broadband emission. In addition to outstanding optical properties, the phosphor demonstrates excellent processability—enabling high-quality films via spin coating and aerosol-assisted chemical vapor deposition (AACVD) from acetonitrile solutions—and rapid, scalable synthesis through gram-scale mechanochemistry. Extending this strategy by replacing TPA+ with methyltriphenylphosphonium (MTPP+) spacers further confirms the structural and luminescent tunability, underscoring the wide-ranging applicability of this approach. Collectively, these results establish the A2[Cu4Br2I4] halide-alloying strategy as a versatile and promising synthetic pathway for developing advanced phosphors suitable for next-generation optoelectronic and lighting technologies.
The reagent quantities for each composition are summarized in Table 1.
| CuBr | CuI | TPA-Br | TPA-I | |
|---|---|---|---|---|
| x | mmol (g) | mmol (g) | mmol (g) | mmol (g) |
| 0 | 2.4 (0.3443) | 0 | 1.2 (0.3195) | 0 |
| 1 | 1.8 (0.2582) | 0.6 (0.1143) | 1.2 (0.3195) | 0 |
| 2 | 1.2 (0.1721) | 1.2 (0.2285) | 1.2 (0.3195) | 0 |
| 3 | 0.6 (0.0861) | 1.8 (0.3428) | 1.2 (0.3195) | 0 |
| 4 | 0 | 2.4 (0.4571) | 1.2 (0.3195) | 0 |
| 5 | 0 | 2.4 (0.4571) | 0.6 (0.1598) | 0.6 (0.1889) |
| 6 | 0 | 2.4 (0.4571) | 0 | 1.2 (0.3759) |
rotoinversion axes. The anion is located on an inversion center and the n glide plane, and therefore is completely disordered. Additionally, the three crystallographically independent X positions all feature disorder between Br and I atoms, where the proportion was refined for each position separately using free variables resulting in a precise formula TPA2[Cu4Br2.16I3.84] for the analyzed crystal. The Cu–X bond lengths are left unrestrained even though they are slightly shorter than the average values reported in the literature, because they represent the statistical average of all the positions, they can have in the anion due to the disorder, and even bonds to two neighboring copper atoms need not be equal. This led to a best fit to the experimental electron density. The hydrogen atoms of the C–H bonds were placed in idealized positions, and refined with Uiso tied to the parent atom. The disordered copper, halogen atoms, and tetrapropylammonium cation were refined with Uij restraints (SIMU, RIGU, EADP) implemented in SHELXL-2019/3. Powder X-ray diffraction (PXRD): patterns were recorded on an Ultima IV Rigaku diffractometer using Cu-Kα radiation (λ = 1.5418 Å) over a 2θ range of 5°–55° with a step size of 0.02°.
Photoluminescence (PL): excitation and emission spectra were recorded on an Edinburgh Instruments FS5 fluorometer equipped with a 150 W ozone-free xenon arc lamp and Czerny–Turner monochromators. Temperature-dependent photoluminescence and decay times were recorded using a cryostat and a 375 ± 5 nm picosecond pulsed diode laser, while photoluminescence quantum yield (PLQY) was measured at room temperature using an integrating sphere on the same fluorimeter. The measurements of the spectral power distribution (SPD) in the wavelength range λ = 400–800 nm of the LED lamp and the converting phosphor layer studied were performed with a CCD spectroradiometer LISUN LMS-9000B with an integrating sphere. Thermogravimetric analysis (TGA) was performed on a TGA Q5000 IR system (TA Instruments, New Castle, DE, USA) under nitrogen flow. Approximately 5 mg of dry sample was heated from 30 °C to 400 °C at a rate of 5 °C min−1.
Single-crystal X-ray diffraction (SCXRD) analysis was performed to determine the atomic structure of TPA2[Cu4Br2I4], revealing that it is isostructural to TPA2[Cu4Br6]. The phosphor crystallizes in the tetragonal crystal system (space group P42/n), with lattice parameters a = b = 15.773 Å and c = 16.055 Å. Its structure comprises four slightly distorted trigonal planar [CuX3]2− subunits, connected by shared edges to form a [Cu4X6]2− cluster (X = Br, I). The asymmetric unit contains half of the [Cu4X6]2− anion and parts of three crystallographically independent TPA+ cations (half, quarter, and quarter), as depicted in Fig. 1c (see Tables S1 and S2 for more crystallographic details).
Within the Cu4 tetrahedron, the Cu–Cu distances are shorter than the sum of the van der Waals radii of Cu(I) (2.80 Å), indicating strong metallophilic interactions.8,22 Furthermore, four iodine and two bromine atoms are disordered across six positions around the cluster. Consequently, two distinct halide configurations are expected depending on bromine positioning: cis, where bromine atoms are adjacent, and trans, where they are alternately positioned (Fig. 1d). The structural integrity of these disordered clusters is preserved by the bulky TPA+ organic spacers. To assess the geometric stability of the cluster, the effective ionic radii of the TPA+ molecules and [Cu4Br2I4]2− clusters were calculated to be 2.545 and 5.314 Å, respectively. Their radius ratio is approximately 0.479, which falls within Pauling's octahedral stability window (0.41–0.73),23 supporting the robustness of the crystal structure.
The phase purity of the obtained solids was corroborated by powder X-ray diffraction (PXRD), which shows an excellent agreement between the experimental PXRD patterns and the pattern calculated from the single crystal structure (Fig. 1e). On the other hand, X-ray photoelectron spectroscopy (XPS) confirms the presence of Cu, Br, and I in the sample (Fig. S1a–c). The high-resolution Cu 2p core level spectrum shown in Fig. 1f displays two binding energy peaks at 952.0 (Cu 2p1/2) and 932.2 eV (2p3/2), consistent with Cu(I) species. Notably, the absence of the Cu(II) related satellite peaks confirms that Cu remains in the +1 oxidation state, which is crucial for the material's stability and optimal luminescence performance.
The presence of excitonic absorption peaks at room temperature suggests a high exciton binding energy, likely supported by the non-conjugated TPA+ spacers that provide a low-polarizability environment.24 This strong exciton confinement is reflected in the photoluminescence behavior, where the emission spectrum exhibits a broad asymmetric band peaking at 526 nm. Additionally, PLQY measurements reveal an impressive 95% efficiency (Fig. 2b), confirming minimal self-absorption despite the low Stokes shift (69 nm). These properties indicate a highly efficient luminescent process, warranting a deeper exploration of the underlying emission mechanism.
To systematically analyze this behavior, PL spectra were collected at 20 K increments and deconvoluted into H and L bands. The normalized temperature-dependent bands are independently plotted in Fig. S2b and c. At 80 K, both bands appear narrowed and seem to be decoupled. Upon heating from 80 to 300 K, the H-band shifts slightly from 518 to 513 nm, with its FWHM increasing from 53 to 66 nm. The L-band undergoes a more pronounced shift, from 613 to 570 nm, with a corresponding FWHM increase from 93 to 110 nm. This blue shift and broadening are characteristic of increased electron–phonon coupling, commonly observed in metal halide luminescent materials.25 At low temperatures, structural rigidity minimizes distortions in the excited state, effectively decoupling excitons from phonons.
In addition, the ratio of integrated H and L bands (H/L) changes from 0.65 at 300 K to 0.99 at 80 K, suggesting that at room temperature, energy transfer between the two states is efficient, while at 80 K, the energy barrier becomes too large to overcome.
Complementary photoluminescence measurements under in situ heating (320–380 K, Fig. S2f) revealed a gradual emission decrease. This process was fully reversible upon cooling, indicating thermally activated quenching rather than permanent structural degradation, consistent with enhanced exciton–phonon interactions at elevated temperatures. Thermogravimetric analysis (Fig. S2g) confirmed high thermal stability, with no significant mass loss up to 105 °C. A minor 2% loss occurred between 105 and 200 °C, followed by linear mass reductions of ∼10% at 255 °C and ∼38% at 320 °C, closely matching the expected mass contribution of tetrapropylammonium bromide in the sample.
At 80 K, the emission mechanisms decouple, as evidenced by distinct double exponential decay times of 4.07 and 11.39 μs. In contrast, at 300 K, the decay profiles become indistinguishable, yielding a monoexponential decay of 3.59 μs. These results, together with the H/L ratio changes, confirm that the energy barriers between excited states are low, enabling excitons to efficiently cross between states and achieve thermal equilibrium. At 300 K, this results in uniform green emission from TPA2[Cu4Br2I4], facilitated by the effective coupling of multiple excited states.
• (x = 0) formation of UV-excitable red-emitting TPA2[Cu4Br6], exhibiting excitation-independent emission from a single excited state.
• (x = 1–3) excitation-dependent emission, indicating the coexistence of [Cu4Br6]2− and [Cu4Br2I4]2− clusters rather than a single [Cu4Br5I]2− species. As iodine content increases, the contribution of [Cu4Br2I4]2− dominates.
• (x = 5) the emergence of a new white-emitting species when excited at 330 nm, alongside TPA2[Cu4Br2I4].
• (x = 6) emission shifts to pure white, corresponding to the formation of the UV-excitable TPA2[Cu2I4], with non-emissive impurities.
The XRD patterns of ground single crystals (Fig. S4) support this evolution, revealing gradual lattice expansion as Br is replaced by I. For x > 4, diffraction peaks characteristic of the zero-dimensional TPA2[Cu2I4] phase become prominent. These results are consistent whit previous reports and emphasize the necessity of precise Br/I balance to achieve the unique luminescent properties of this phosphor.12
In stark contrast, TPA+ lacks any π-conjugated system, rendering it optoelectronically inert and incapable of participating in MLCT or HLCT processes. Consequently, the conventional MLCT/HLCT-based explanation for the two emission bands cannot account for the photophysical behavior observed in TPA2[Cu4Br2I4].
TD-DFT has been extensively employed to investigate copper-based systems, consistently achieving strong agreement with experimental data by accurately describing charge-transfer phenomena and excited-state geometric distortions.19,29–32 Specifically, the tHCTHhyb functional has proven effective in reproducing key excited-state properties, including absorption and emission energies, as well as the accurate depiction of triplet states.33 Building on this foundation, we conducted an electronic structure analysis of the [Cu4Br2I4]2− clusters in both cis and trans configurations, using the optimized tHCTHhyb functional. This analysis, leveraging single-crystal structural data (Fig. 3), enabled a detailed investigation of the ground and excited states, providing critical insights into the origins of the high (H) and low (L) energy emission bands.
Notably, the cis and trans structures are virtually isoenergetic in the ground state (ΔE = 0.1 kcal mol−1), consistent with the observed disorder in the SCXRD analysis. The excited-state dynamics are represented using configurational coordinate diagrams (Fig. 3a and b). Vertical photoinduced S0 → S1 transitions for the trans and cis structures occur at 3.12 eV and 3.09 eV, respectively—values slightly higher than the experimental bandgap (2.8 eV), but within reasonable theoretical deviation.
From the S1 state, both structures engage in a highly efficient intersystem crossing (ISC) cascade into the 3CC triplet states, initiating from T6 and T7, for the cis and trans configurations, respectively (see Fig. S5). These triplet states sequentially relax internally to the T1 state, which subsequently deactivates radiatively to S0. This multi-step relaxation pathway, characteristic of triplet-state dynamics, directly accounts for the long photoluminescence (PL) decay times observed experimentally. For the trans configuration, the energy gap between T1 and a high-energy S0 conformer is 2.36 eV, followed by a final relaxation of 0.31 eV to the ground-state minimum. In contrast, the cis configuration exhibits a smaller T1 to S0 energy gap of 1.79 eV, accompanied by a more pronounced structural rearrangement of 0.63 eV upon reaching the ground state.
The calculated emission energies from these relaxed T1 states align well with the experimentally observed H and L emission bands, measured at 2.4 eV and 2.02 eV, respectively. Moreover, the calculated exciton energies—defined by the energy difference between the optimized S0 and T1 states—are 525.36 nm for the trans structure and 692.65 nm for the cis structure. This theoretical behavior is consistent with the experimental trend observed as temperature increases, where the two emission bands undergo a progressive blue shift.
The strong agreement between theory and experiment validates our model, allowing for a confident interpretation of the photoinduced electronic redistribution and the associated structural changes governing the dual emission behavior.
In our previous work,34 we established a mechanism where photoexcitation induces electron density redistribution within the valence shell of copper atoms, triggering subsequent geometric rearrangements. This mechanism is directly applicable to the [Cu4Br2I4]2− isomers, where the structural changes observed can be attributed to photoinduced electron transfer between atoms within the cluster. Specifically, in the trans configuration, a redistribution of 0.16 e is observed, where two copper atoms and four iodide atoms donate electron density to the remaining two copper atoms and two bromide atoms. In the cis configuration, three copper and three iodine atoms transfer 0.28 e to the rest of the cluster.
These photoinduced electron redistributions result in distinct structural distortions during the excited-state relaxation process, which in turn directly influence the dual emission behavior. The connection between these theoretical predictions and the experimentally observed high (H) and low (L) energy emission bands is further reinforced by the calculated exciton energies and structural dynamics.
As illustrated in Fig. 3c and d, the structural changes that occur between the S0 and T1 states differ significantly between the two isomers. For the cis configuration (Fig. 3c), the most prominent geometric change is an expansion of the Br–Cu–Br angle from 119.05° to 131.3°, coupled with a bromide ion shifting 0.5 Å away from the cluster. This angular distortion, which stabilizes the T1 state, makes the cis configuration more susceptible to electron–phonon coupling, directly contributing to the pronounced temperature dependence of the L emission band energy.
Conversely, in the trans configuration (Fig. 3d), the primary structural change is an increase in the Br⋯Br distance by 0.29 Å and a reduction in Cu⋯Cu distances, with the most significant contraction of 0.31 Å between the copper atoms that lose electron density. This structural reorganization is consistent with the energy stabilization of the T1 state, leading to the observed H emission band.
The strong correlation between our theoretical analysis and the experimental observations—including the temperature-dependent redshift of the two emission bands and the long-lived triplet state dynamics—provides compelling support for the proposed model. These findings establish a clear connection between photoinduced structural reorganization and dual emission behavior, not only deepening our understanding of this system but also offering a robust framework for designing copper-based luminescent materials with tunable properties.
Versatile deposition methods underscore the practical applicability of this phosphor. Films prepared via spin coating on top of glass substrate (Fig. 4a and b), and aerosol-assisted chemical vapor deposition (AACVD) on a flexible PET sheet (Fig. 4c), successfully retain the luminescent properties. SEM analysis (Fig. S7a and b) confirms good substrate coverage, revealing distinct morphological signatures: spin-coated films exhibit rod-shaped grains (∼0.3 × 1 μm) coalescing in island-like formations (characteristic of centrifugal-driven crystallization); whereas AACVD-produced films feature well-packed, faceted grains (5–8 μm) resembing the tetragonal crystalline morphotype. AACVD, in particular, stands out for its versatility, scalability, and adaptability to diverse device architectures including non-conventional flexible substrates.
To illustrate practical potential, composite films incorporating mechanochemically synthesized phosphor within a polymer matrix were successfully employed as wavelength-converting phosphors atop a commercial high-power royal-blue LED (10 W), efficiently generating warm-white emission (Fig. 4d and e). Spectral details are provided in Table S3, emphasizing the material's suitability for practical optoelectronic applications.
To highlight the simplicity, versatility, and broad applicability of our halide alloying strategy, we substituted the TPA+ spacer with methyltriphenylphosphonium (MTPP+). This phosphor was synthesized via the rapid mechanochemical route, similarly achieving tunable and highly promising luminescent properties. Photoluminescence excitation and emission spectra of the resulting MTPP2[Cu4Br2I4] phosphor (Fig. 5a) show excitation-independent emission, with no spectral shifts observed upon varying the excitation wavelength, analogous to the TPA-based system. Compared to the TPA analogue, the MTPP emission maximum is slightly red-shifted (545 nm, Fig. S8a) and requires a more complex spectral deconvolution, suggesting multiple contributing processes that may include MLCT/HLCT in addition to cis/trans configuration-related emission (Fig. S8b). For comparison, spectra of the previously reported iodide-only MTPP2[Cu4I6] system are included, further highlighting the effectiveness of halide alloying in precisely modulating emission characteristics. The PLQY of the alloyed system was measured to be 72%, comparable to the MTPP2[Cu4I6] composition.8
Powder XRD analysis of MTPP2[Cu4Br2I4] reveals excellent structural agreement with simulated patterns from MTPP2[Cu4I6], exhibiting diffraction peak shifts due to lattice contraction upon bromide substitution (Fig. 5b). These results underscore the generality and power of this alloying approach, enabling targeted control over luminescence properties across diverse organic spacers employed in luminescent [Cu4X6]2− clusters.
Collectively, these findings illustrate the exceptional versatility and technological promise of A2[Cu4Br6−xIx] phosphors. This approach could be extended to other UV-excitable A2[Cu4Br6] structures reported with A-cations such as ethyltriphenylphosphonium (ETPP+),35 tetrabutylphosphonium (TBP+),36 and tetraethylphosphonium (TEP+).26 Furthermore, it may provide a strategy to activate systems that are otherwise non-emissive at room temperature, such as the FA2[Cu4Br6] phase,37 where longer A-site-Cu+ distances structurally favor non-radiative decay. In such cases, halide alloying could modify the cluster environment, stabilize excited states, and promote radiative recombination. The demonstrated simplicity, scalability, and versatility in processing, open exciting practical pathways for commercial applications, such as large-scale lighting systems and flexible optoelectronic devices. Current ongoing investigations, including optimized growth of large centimeter-scale single crystals, advanced polymer composite development, refinement of spin-coating techniques for low-scattering solar concentrators, and versatile AACVD film deposition onto unconventional substrates (textured, flexible, elastic), promise significant extensions of the practical applicability of these phosphors, paving the way toward next-generation luminescent and optoelectronic technologies.
The phosphor's processability was validated through rapid mechanochemical synthesis scalable to industrial production, alongside successful thin-film fabrication via spin coating and AACVD on conventional and flexible substrates. Demonstrated integration into polymer composites and subsequent warm-white LED application further underscores practical relevance. Importantly, this alloying strategy proves transferable to diverse organic spacers, significantly broadening its applicability across various luminescent platforms.
The authors confirm that the data supporting the findings of this study are available within the article and its supplementary information (SI). Supplementary information: characterization data of new compounds. See DOI: https://doi.org/10.1039/d5tc02475g.
| This journal is © The Royal Society of Chemistry 2025 |