Open Access Article
Biljana
Pejova
*a,
Julio A.
do Nascimento
b,
Fayzah
Talbi
bc,
Thomas Jack
Fawcett-Houghton
b,
Adam
Kerrigan
be,
Leonardo
Lari
be,
Richard E.
Douthwaite
d,
Ljupcho
Pejov
a and
Vlado K.
Lazarov
*be
aInstitute of Chemistry, Faculty of Natural Sciences and Mathematics, SS. Cyril and Methodius University, POB 162, 1000 Skopje, North Macedonia. E-mail: biljana@pmf.ukim.mk
bSchool of Physics Engineering and Technology, University of York, York, YO10 5DD, UK. E-mail: vlado.lazarov@york.ac.uk
cDepartment of Physical Sciences, Faculty of Science, University of Jeddah, P.O. Box 13151, Jeddah, 21493, Saudi Arabia
dDepartment of Chemistry, University of York, York, YO10 5DD, UK
eThe York-JEOL Nanocentre, University of York, York, YO10 5DD, UK
First published on 12th July 2025
In this work, we demonstrate the colloidal bottom-up synthesis of spinel AgIn5S8 quantum dots (QDs) with tunable optical properties. The QD size, and consequently their band gap energy (Eg), is effectively controlled by reaction temperature and ultrasound (US) irradiation. Under combined conditions of 75 °C and US irradiation, ultrasmall QDs with an average size of 2.6 nm are obtained, exhibiting a wide band gap of 3.77 eV. In the absence of US, reactions conducted at 55 °C and 75 °C yield larger QDs (∼5 nm and 31 nm, respectively), with reduced band gaps of 3.09 eV and 2.18 eV. The elevated temperature (75 °C) suppresses sulfur-chain formation that otherwise limits growth at 55 °C, while acoustic cavitation induced by US enables narrowest size distribution. Annealing of as prepared QDs, at 200 °C for 2 h, promotes coalescence resulting in QDs with increased size of ∼34 nm, with a bulk like band gap of 1.73 eV for QDs prepared without US. In contrast, annealing of the QDs, prepared with US, results in polycrystalline QDs with average size of ∼21 nm. High-resolution transmission electron microscopy reveals a strong correlation between QD size, structural ordering and optical behavior. The as-prepared 2.6 nm QDs exhibit lower Urbach energy, attributed to their single-crystalline nature, unlike the less ordered QDs synthesized without US. Annealing improves structural ordering and reduces Urbach energy in QDs prepared at 75 °C, while stacking faults and grain boundaries in other QDs hinder such improvements. Photoluminescence measurements further confirm a strong relationship between QD structure, size, and emission characteristics. The synthesized AgIn5S8 QDs exhibit remarkable band gap tunability of up to 2 eV across the visible spectrum and sharp band-edge emission, underscoring their potential for applications in optoelectronic and biomedical devices. This work provides a robust and sustainable pathway to high-performance, non-toxic QDs, addressing a key bottleneck for their use in biocompatible and consumer electronics.
The quantum confinement effect allows the band gap of QDs to be precisely tuned by controlling their size and composition.6–9 Their strong luminescence, narrow emission bandwidth and excellent photostability, combined with low toxicity in selected systems, make them promising candidates also for bioimaging, biosensing and diagnostic applications.10–12 Particularly, ultrasmall QDs (<3 nm) are of great interest for biomedical use due to their ability to cross biological barriers and avoid immune recognition, mimicking the behavior of small-molecule drugs.13,14
To date, most QD research has focused on binary systems such as II–VI, III–V and IV–VI semiconductors.15–19 However, these materials often incorporate toxic heavy metals (e.g. Cd, Pb, Hg), posing challenges for environmental and biomedical applications. This has motivated a shift towards less toxic alternatives such as ternary I–III–VI semiconductors (I = Cu, Ag; III = In, Ga, Al; VI = S, Se, Te).20–26 These chalcogenide semiconductors, known for their broad structural and compositional tunability, offer direct band gaps in the visible range and high absorption coefficients (104–105 cm−1).12 Their properties can be further engineered through doping and phase control.
Among these, Ag–In–S based QDs are particularly attractive due to their versatile photoluminescence (PL) characteristics. Compared to other I–III–VI systems, AgInS2 QDs exhibit smaller band gaps due to the influence of In3+ and S2−, resulting in red to near-infrared emission.27 Doping with smaller cations (e.g., Ga3+, Al3+, Bi3+, Zn2+) can shift the PL toward shorter wavelengths (green or blue), offering tunable emission across the visible spectrum.28–30 As such, Ag–In–S QDs are emerging as a non-toxic alternative for applications including biocompatible imaging, photocatalytic pollutant degradation and solar hydrogen production.31–62 However, most studies to date have focused on AgInS2, with limited work on other stoichiometries such as the spinel-phase AgIn5S8.
In this work, we focus on AgIn5S8, a direct-band-gap semiconductor with high PL quantum yield, energy convergence in the excited state, desirable for quantum light-emitting diode applications.27,49,50 We demonstrate a low-temperature, aqueous-phase colloidal synthesis method – both with and without ultrasound assistance – for the preparation of AgIn5S8 QDs with controlled size and optical properties.
Unlike traditional syntheses in organic solvents which often require complex procedures and post-synthetic modifications (e.g. ligand exchange, silica coating) to achieve water dispersibility,53 our aqueous-based approach directly yields water-dispersible QDs under mild conditions. Organic-solvent-based routes also tend to suffer from poor reproducibility and challenging scale-up. Importantly, aqueous synthesis of phase-pure AgIn5S8 QDs has been hindered by competing formation of secondary phases such as AgInS2, In2S3, and Ag2S.54,55 Our method circumvents these limitations, producing phase-pure spinel AgIn5S8 QDs with exceptional band gap tunability.
We report the broadest band gap tuning range for any Ag-based ternary QD system to date, from 1.70 to 3.77 eV. By comparison, AgInS2 QDs typically exhibit tunability only between 2.3 and 3.1 eV.56 Moreover, previous attempts to achieve narrow, band-edge photoluminescence in I–III–VI QDs have been limited. Time-resolved PL and X-ray studies have indicated that broad emission in these systems originates from multiple donor–acceptor recombination pathways resulting in PL full width at half maxima (FWHM) exceeding 100–300 meV.57–59 In contrast, our AgIn5S8 QDs exhibit significantly narrower PL linewidths: FWHM as low as ∼25 meV in annealed samples and ∼100 meV in as-deposited nanocrystals. These values are within or even better than industrial standards and notably better to those of typical II–VI QDs such as CdSe, CdTe and InP.63,64
Overall, this work advances the field by providing a green, scalable and reproducible synthesis of high-performance spinel Ag-based QDs, aligning with sustainable chemistry goals through low-energy processing and use of water as a safe solvent.
The main components of the reaction system are ionic precursors and a ligand agent. AgNO3 and In2(SO4)3 were used as cation (Ag+ and In3+) precursors, while thioacetamide, (CH3)CS(NH2) is the sulfide precursor. Sodium thiosulfate, Na2S2O3 was included in the reaction system as a ligand agent to control the concentrations of the relevant ions. This approach requires carefully controlled equilibria during the reaction process in the aqueous solution which are mutually and pH dependent, see ESI-1† for more details.
Synthesis of AgIn5S8 was achieved in all three approaches. We note that our aqueous medium method of AgIn5S8 QDs synthesis is greener compared to more severe organic-based synthesis. Furthermore, multi-bubble sonoluminescence (MBSL) conditions, in case of US-assisted route, are also quite mild65 in comparison with the other typically high temperature wet chemical methods.
Of note during synthesis color changes were observed for all reactions. For QD-75US the reaction system went from transparent through greenish, yellow and orange to red (ESI Fig. 1†) with sonoluminescence65,69,70 observed during the reaction and a red-shift of the emitted colors suggesting QD growth during synthesis. The formed QDs absorb energy, released by acoustic cavitation and emit bright colors depending on their size-dependent energy spectrum. Similarly, for QD-55 and QD-75, chemiluminescence phenomenon is observed, with redshift of color change again indicating growth of QDs during the reaction process.
XRD patterns were measured on an Ultima IV X-ray diffractometer (Rigaku Co.) using CuKα radiation in the 2θ/θ scanning mode and step-scan of Δ(θ) = 0.02°. Using experimentally obtained XRD patterns, the average size of QDs was calculated using Scherrer's approach. The (2θ) and full-width at half maxima (β) of XRD peaks have been determined by fitting with linear combination of Gaussian and Lorentz functions. The instrumental contribution to the broadening of XRD peaks has been taken into account applying variance subtraction approach.
SEM images were collected using a JEOL 7800F Prime SEM. TEM selected area electron diffraction (SAED) patterns, as well as high resolution TEM (HRTEM) images were obtained using a JEOL 2100+ TEM and aberration corrected JEOL-NeoARM STEM/TEM. Diffraction patterns have been simulated using JEMS software and the TEM-SAD patterns were calibrated using Au nanoparticles as a reference. The atomic interplanar spacings were measured from selected areas of HRTEM images using fast-Fourier transform (FFT).
Spectral dependences of transmittance were measured in the spectral range from 300 nm to 1100 nm on a Cary 50 spectrophotometer (HP Co.).
Steady-state photoluminescence (PL) spectra of all the synthesized samples were recorded at room temperature on a Hitachi F4500 fluorescence spectrophotometer with a 150 W Xe lamp as a source of excitation, a diffraction grating monochromator and a photomultiplier tube detector. Excitation-emission 3D matrices (EEMs) were measured in reflection geometry. The corresponding photoluminescence excitation spectra (PLE) at relevant emission wavelengths have also been measured. PL spectra were recorded for both solid-state samples as well as NP dispersions in ethanol.
The Raman spectra were measured on a LabRam300 spectrometer by Horiba Jobin-Yvon, using the red 632.8 nm He:Ne laser line and a backscattering mode.
Structural determination by XRD and SAED of QD-55, QD-75 and QD-75US show that all synthesized particles have the spinel structure. Fig. 1a shows the reference peaks of the spinel phase AgIn5S8, which are well matched with the QD-75, showing similar peak intensity ratios indicating that QD-75 have spherical-like shape. In contrast, the QD-75US and QD-55 XRD peaks (Fig. 1c and d), are significantly broadened around the most intense (311) and (440) peaks which is due to the significant size reduction of these QDs.
![]() | ||
| Fig. 1 XRD patterns of: (a) reference peaks for AgIn5S8, (b) QD-75, (c) QD-55 and (d) QD-75US. Reference peaks for AgIn5S8 are taken from ref. 71. The peak, appearing in (c), labeled by * is due to the polymeric sulfur.72 | ||
Furthermore, the SAED from all QDs presented in Fig. 2 clearly show all diffraction rings associated with AgIn5S8 spinel phase, as seen by the fits of the calculated SAED with the experimental rings.
![]() | ||
| Fig. 2 SAED patterns of QD-55 (a), QD-75 (b) and QD-75US (c) with diffraction rings corresponding to AgIn5S8. The insets are BF-TEM images from the areas from which the SAED pattern were taken. | ||
Careful analysis of the XRD (311) peak from QD-55 and QD-75US shows that there is a subtle difference in their shape (Fig. 1c and d), indicating that QD-75US has two distinct populations, as seen in Fig. 3. The larger QD-75US contribute to the narrow tip (Fig. 3a), while the population of smaller QD-75US contribute to broader tails of the (311) peak. On the other hand, a good fit of this peak for QD-55 has been obtained without a contribution of a narrow function (Fig. 3b). Two populations of QD-75US are also visible in the (440) XRD peak and, as expected, only one population for QD-55, as shown in the supplementary material, ESI Fig. 2a and b.† Utilizing the Scherrer approach, the calculated value of QD-75 average diameter is 31.2 nm. Details of the calculations are shown in ESI (ESI Fig. 3†). The average diameter of larger population of QD-75US, calculated on the basis of (311) and (440) peaks, is 19.2. Smaller QD-75US and QD-55 have average diameter less than 5 nm.
In order to confirm the QD size obtained by the Scherrer method, and further to study the micro and atomic structure of the QDs, TEM/STEM was performed. Fig. 4 shows low magnification images of the QDs, where the overall spherical shape of the aggregated QDs is observed. As confirmed by further HRTEM measurements, however, the aggregates are composed of distinct QDs.
The size distribution obtained by measuring their diameter, from TEM images, is shown in Fig. 5, and we observe a good agreement between average QDs size obtained by TEM and XRD. Performing a fitting procedure using a lognormal function, gives (4.9 ± 2.4) nm and (31.1 ± 9.0) nm for the average diameter of QD-55 and QD-75, respectively. For QD-75US the small and large populations are (2.6 ± 0.7) nm and (21.0 ± 3.3) nm, respectively.
Further study by TEM of QD-55 reveals fiber-like structures (Fig. 6). The fibers are due to the formation of polymeric sulfur, which are indicated in Fig. 6a and b as well as surrounding AgIn5S8 QDs. Polymeric sulfur72 is observed in XRD measurement (denoted with * in Fig. 1c) and using Raman spectroscopy shown in ESI Fig. 4a.† Although the peak due to polymeric sulfur seems to appear very close to the (220) reflection of AgIn5S8 QDs in the XRD pattern in Fig. 1c, a closer examination of the corresponding region clearly reveals that its position is shifted to lower 2θ values corresponding to polymeric sulfur (ESI Fig. 4b†). Comparison with literature data reveals the Raman bands characteristic for polymeric sulfur denoted with asterisks (ESI Fig. 4b†). In addition, presence of shoulder band (denoted with red arrow) can be assigned to accretion of polymeric sulfur chains (Sμ), which indicates a certain level of disorder.73 Note that such spectral features are absent in the Raman spectra of QD-75 and QD-75US. Atomic spacing measurements by intensity profile of HRTEM images of sulfur chains show the atom-to-atom measurements in this fiber-like structures, ∼3.5 Å, that correspond to reported values of sulfur polymeric chains, Fig. 6. The excess of sulfur on reaction at 55 °C, implies that the reactants are not fully converted/consumed into the aimed final AgIn5S8 product.
The fiber structures, also illustrated by STEM imaging of QD-55 (ESI Fig. 5†), are absent for QD-75 (Fig. 7). This demonstrates that the reaction temperature significantly influences the mechanism of QD formation, resulting in formation of large QDs with average diameter of over 5 times larger than the QD-55. The HRTEM analysis of QD-75 shows that the majority have a single crystal structure (Fig. 7), which becomes polycrystalline at larger sizes, Fig. 7a. This suggests that large particles in QD-75 are likely obtained by coalescence of smaller NPs which leads to the formation of structural domains, as observed in Fig. 7a. The influence of these structural boundary defects in the QDs will influence their optical properties, as will be discussed below.
![]() | ||
| Fig. 7 (a) HRTEM of QD-75, an enlarged QD-75 is shown in (b). (c) Polycrystalline QD-75, clearly showing the structural domains and grain boundaries within a single QD-75. | ||
The above experimental constraints that limit controlling QDs size can be overcome using ultrasound (US) assisted synthesis. The application of US leads to free radicals such as H˙ and HO˙ when the cavitation medium is water (due to water homolysis), as well as reactive oxygen species which influence the precursor conversion and enhance nucleation rate.74–76 The implosion of bubbles is nearly adiabatic and produces temperature in the thousand-degree range in the microbubbles for a very short time. This effect can be rather effective for modifying structural properties of QDs via local annealing effects since AgIn5S8 is characterized by relatively low melting temperature (∼ 1110 °C (ref. 77)). While energy released as a consequence of the collision between QDs can invoke melting of AgIn5S8 and QD enlargement, acoustic cavitation followed by a variety of physical effects, such as shock waves and micro-jets, it is expected to have a dominant effect, providing a potential pathway for QD size reduction.
Indeed, the use of US enables a QD diameter reduction from (31.1 ± 9.0) nm to (2.6 ± 0.7) nm, confirming that the heterogeneous acoustic cavitation effects are dominant for QD-75US.74,75 Furthermore, the additional minor subpopulation of QD-75US with larger size can be attributed to coalescence effects. Synthesis using US conditions also affects the crystallinity and defects of QD-75US. HRTEM shows that the majority and minority of QD-75US population histograms, shown in Fig. 5, have average diameter sizes of (2.6 ± 0.7) nm and (21.0 ± 3.3) nm respectively, with typical QDs shown in Fig. 8. The HRTEM imaging of ultrasmall QD-75US show that these NPs have single crystal structure, though even at this size range occasionally QD-75US with extended defects such as twins can be observed in ESI Fig. 6.† In contrast to QD-75, the larger population of QD-75US do not show single crystal structure, instead they show crystallographic domains (ESI Fig. 7†), indicating that they are formed via coalescence of smaller QDs.
![]() | (1) |
![]() | ||
| Fig. 9 (αhv)2 on hv dependence of QD-75 (blue), QD-55 (orange) and (c) QD-75US (green). The intercepts with (hv) axis give the Eg values of the QDs. | ||
The >10 times reduction of the average QD diameter for QD-75US in comparison to QD-75 results in blue-shifted band gap energy for ∼2 eV when compared to the bulk band gap of AgIn5S8. We note that Eg of bulk AgIn5S8 reported in the literature varies in the range between 1.7 to 1.9 eV.78–81 Although the Eg of QD-75 is significantly red shifted with respect to QD-75US, it is still slightly blue-shifted by ∼0.4 eV compared to bulk AgIn5S8, showing that quantum confinement effects are observed even for this set of particles with an average diameter of (31.1 ± 9.0) nm. We note that the theoretical study82 of Eg confirms that the valence band maximum (VBM) and the conduction band minimum (CBM) correspond to the same point of k-space (Γ: 0.0, 0.0, 0.0), the calculated band gap value for the modelled AgIn5S8 compound (Eg = 1.2 eV) is significantly lower than the experimentally obtained values. This discrepancy is partially due to the structure of the QDs spinel phase where there is need to consider the random distribution of the 1/5 of indium ions over the tetrahedral sites along with silver ions, as well as inherent systematic errors when DFT approach is applied to excited states.
The long tail in the UV-Vis spectrum below the band gap is due to exponential sub-band gap absorption,83–85Fig. 11. The data from this region can be analyzed according to the Urbach–Martienssen rule:83
![]() | (2) |
![]() | (3) |
The values of Urbach energy, obtained from the slope of lnα on photon energy dependences of QD-75US, QD-55 and QD-75 (Fig. 10), are (916 ± 56) meV, (1097 ± 37) meV and (1271 ± 22) meV respectively. The smallest Urbach energy value of QD-75US implies that the US-assisted route leads towards a lower degree of inherent structural disorder compared to the QDs synthesized under purely thermal conditions. These data correlate with HRTEM analysis (Fig. 7), where the structure of QD-75 is less-ordered compared to the QD-75US, clearly showing structural domains separated by the grain boundaries that deviate from ideal bulk atomic stacking. It is also expected that the minor population of large QD-75US would also affect the sub-band gap absorption albeit below the region of Urbach absorption of the smaller QDs, due to its lower Eg.
The different band gap (Eg) and Urbach (Eu) energies arising from size and structural differences are expected to significantly influence the PL properties of the AgIn5S8 QDs and have been probed using excitation-emission steady-state photoluminescence (PL). The relevant regions of the PL spectra of the QDs are shown in Fig. 11 as contour plots. Upon elimination of the first-, second- and third-order Rayleigh scattering and Raman scattering features, the persistent PL features were analyzed and assigned. In this section, the focus is on emission bands due to band-to-band electronic transitions.
Fig. 11a shows the PL spectrum of QD-75 obtained as a slice through the 3D plot at an excitation wavelength of 500 nm. The PL spectrum is dominated by an intense emission band at ∼610 nm (2.03 eV) and a satellite band at ∼590 nm (2.10 eV), see ESI Fig. 8a.† The transition energies of both bands lie slightly lower in energy than the band gap energy determined from absorption spectroscopy. We attribute this finding to the adiabatic optical absorption–induced interband electronic transition which leads directly to an excited vibrational state(s) within the excited electronic state, according to the Franck–Condon principle. The corresponding interband transition due to emission, on the other hand, starts from the ground vibrational state of the excited electronic state of the QD and leads to an excited vibrational state within the ground electronic state and thus occurs at lower energy than the transition due to optical absorption.
Fig. 11b, at an excitation wavelength of 280 nm of QD-55, shows two dominant Gaussian-type emission bands at ∼420 nm (2.95 eV) and ∼440 nm (2.82 eV). Both transition energies are, again, somewhat below the optical band gap energy of 3.1 eV. We attribute the higher – energy (2.95 eV) PL band to the band-to-band recombination emission. The lower energy band (2.82 eV), with higher PL intensity, is attributed to the recombination involving shallow band gap states located near the conduction or valence band edges.
Fig. 11c shows the PL spectrum of QD-75US obtained at excitation wavelength of 280 nm. The centroid of the lowest-wavelength emission band in this case appears at ∼350 nm (3.54 eV), slightly below the optical band gap value determined from analysis of the absorption spectroscopy data (3.77 eV). This band is notably structured from the higher-wavelength side, although these features are intensity-amplified by the band due to second-order Rayleigh scattering band appearing at doubled excitation frequency (560 nm).
The static PL data (concerning band-to-band emission wavelengths) are in excellent correlation with the band gap energies corresponding to fundamental band-to-band transitions computed from the optical absorption spectra. Fig. 12 shows the superimposed optical absorption and photoluminescence spectra of QD-55 and QD-75US. Positions of the PL band maxima are plotted vs. the corresponding optical band gap energies in ESI Fig. 8b.† Particularly, the shifts in band-to-band emission energies in the series QD-75, QD-55 and QD-75US, obtained from the static PL spectra are in excellent agreement with the shifts in optical band gap energies, which further confirms the correct assignments and claims outlined before in the manuscript (ESI – Table 1†). It is important to note that the half-widths of the PL bands are about 100 meV in the case of QD-75. As shown in the next chapter, they are significantly narrowed upon thermal annealing treatment. These values are, however, already comparable to those which are considered as industry standards.63,64
![]() | ||
| Fig. 13 Spectral dependences of absorption coefficient (a) QD-55, (b) QD-75 and (c) QD-75US of unannealed QDs (blue) and annealed QDs at 200 °C for 2 h (orange), respectively. | ||
The red-shifted band gap energy (Fig. 14) indicates QDs enlargement due to sintering upon annealing, as well as different population sizes before and after thermal treatment. The increase in QD size and overall crystallinity has been confirmed by XRD (ESI Fig. 9†) and TEM measurements. The XRD peaks become narrower and clearly defined particularly in comparison to the smaller QD-55 and QD-75US. The average size of all QDs increased with QD-55 and QD-75 giving diameters of 33.5 nm and 33.9 nm respectively, while QD-75US gives 23.3 nm (ESI Fig. 10†). While convergence of the Eg of all QDs towards the bulk band gap value of AgIn5S8 is found, the Urbach energies significantly decreased only for QD-75, from (1271 ± 22) meV to (855 ± 5) meV (Fig. 14 and Table 1).
![]() | ||
| Fig. 14 (αhv)2vs. hv dependence of (a) QD-75, (b) QD-55 and (c) QD-75US unannealed (blue) and annealed at 200 °C for 2 h (orange), respectively. | ||
| E g/eV | E u/meV | ||
|---|---|---|---|
| QD-55 | Unannealed | 3.09 ± 0.02 | 1097 ± 37 |
| Annealed | 1.74 ± 0.01 | 1023 ± 19 | |
| QD-75 | Unannealed | 2.18 ± 0.01 | 1271 ± 22 |
| Annealed | 1.73 ± 0.01 | 855 ± 5 | |
| QD-75US | Unannealed | 3.77 ± 0.01 | 916 ± 56 |
| Annealed | 1.74 ± 0.02 | 990 ± 44 |
In contrast the Urbach energy for QD-55 and QD-75US remains similar (Fig. 15). The experimentally found Urbach energies suggest different levels of structural/chemical ordering between the QD-75 and QD-55/QD-75US. As seen from the table the Urbach energies for QD-55 and QD-75US remain the same within the experimental error before and after the annealing, indicating less ordered structure of these QDs after the annealing process in comparison to QD-75.
Indeed, a difference in structural ordering has been observed by HRTEM after annealing (Fig. 16). While overall improvement of QD crystallinity is observed, QD-55 shows an increase in stacking faults between the large crystalline grains, as well as multiple grain boundaries within single QDs (Fig. 16a). These extended defects are less present in annealed QD-75US, and we note also the presence of a population of small QD-75US (Fig. 16c), which likely is the main contributor of overall structural disorder in these QDs. In the case of QD-75, the number of grain boundaries and stacking faults is decreased, though occasionally twin defects and other type of defects are present (Fig. 16b). The structural analysis of annealed QDs show that there is correlation between the ordering and initial QD size, implying that annealing of small QDs population drives coalescence and overall improvement of crystallinity. However, for QD-55 and QD-75US the presence of extended defects is the main cause of unchanged Urbach energies before and after annealing at the chosen conditions. In contrast, the large size of QD-75 and relatively good crystallinity before annealing (Fig. 1b), shows that the annealing process helps to improve structural/chemical ordering (with only small change in size of ∼3 nm), reflected in a change of the band gap of 0.4 eV and decrease of the Urbach energy of ∼400 meV.
We note that additional annealing of QD-55 and QD-75 for 2 h was not followed by further decrease of band gap energy, the bulk value of Eg has been already reached upon annealing at the mentioned temperature for 2 h (Table 1).
It is important to mention in this context that I–III–VI QDs have been also shown to exhibit notable tunability of the band gap energy upon doping, too (see, e.g. ref. 86).
The effect of post-synthetic annealing on the PL properties can be seen from Fig. 17 where excitation wavelength of 520 nm for annealed QD-75 shows a single emission band at ∼780 nm (1.59 eV), as compared to the band gap energy calculated from optical spectroscopy data of 1.73 eV. This band is evidently much less structured than the bands observed in the as synthesized QD-75 attributable to the more uniform QDs size distribution in this case. Fig. 18 shows the superimposed optical absorption and photoluminescence spectra of as-deposited and annealed QD-75, respectively. The decrease of FWHM of the PL band is evident, the final value of this parameter converging to ∼25 meV in the case of QD-75 annealed samples. To summarize the results concerning this aspect, in stark contrast to previous reports on I–III–VI nanomaterials, our QDs exhibit sharp, band-edge emission, with PL band halfwidths which are within or better than industrial standards, and notably better than those of typical II–VI QDs such as CdSe, CdTe, and InP.63,64 This clearly indicates high compatibility of the presently synthesized material with the industry standards.
![]() | ||
| Fig. 18 Superimposed optical absorption and photoluminescence spectra of QD-75, unannealed (a) and annealed (b). | ||
Thermal annealing of the QDs synthesizes at 55 and 75 °C (QD-55 and QD-75), for 2 h at 200 °C, increases their size to ∼34 nm, while the QD-75US reaches a smaller post-annealing size of ∼23 nm. All annealed QDs exhibit a band gap energy approaching the bulk value of ∼1.7 eV.
The achieved size range—from 2.6 nm to 34 nm—enables an unprecedented degree of band gap tunability for a ternary Ag-based system, spanning a range from 1.70 eV to 3.77 eV i.e. approximately 2 eV across the visible spectrum, demonstrating the remarkable potential of the AgIn5S8 system.
Photoluminescence (PL) measurements reveal a strong correlation between emission characteristics and QD structure and size. Crucially, in stark contrast to previous reports on I–III–VI nanomaterials, our QDs exhibit sharp, band-edge emission. The FWHM is as narrow as ∼25 meV for annealed samples, a value that is several times lower than that of industry-standard II-VI QDs (e.g., CdSe, InP) and well within the stringent requirements for high-definition displays.
While XRD and electron diffraction confirm the cubic spinel structure in all cases, HRTEM uncovers varying degrees of structural ordering. Larger QDs formed via coalescence exhibit multiple structural domains and grain boundaries, contributing to increased Urbach energies in as-prepared samples. Annealing improves structural ordering most significantly in QD-75, where initial and post-annealing sizes are similar. In contrast, QD-55 and QD-75US undergo considerable coalescence during annealing, forming new grain boundaries and stacking faults which essentially do not change their Urbach energies.
Overall, this work highlights AgIn5S8 QDs as a highly versatile material platform with exceptional band gap tunability. Further improvements in photophysical properties are anticipated through strategies such as doping and surface passivation. These enhancements, combined with the inherent biocompatibility and non-toxicity of AgIn5S8, position this system as a strong candidate for optoelectronic and biomedical applications. This study therefore provides key insights into structure–property relationships and low-cost wet-chemical strategies for AgIn5S8 QD synthesis.
Footnote |
| † Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5nr01665g |
| This journal is © The Royal Society of Chemistry 2025 |