Next Article in Journal
Anisotropic Elasticity, Spin–Orbit Coupling, and Topological Properties of ZrTe2 and NiTe2: A Comparative Study for Spintronic and Nanoscale Applications
Previous Article in Journal
A Low-Cost Electrochemical Cell Sensor Based on MWCNT-COOH/α-Fe2O3 for Toxicity Detection of Drinking Water Disinfection Byproducts
Previous Article in Special Issue
Fiber Optic Micro-Hole Salinity Sensor Based on Femtosecond Laser Processing
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

AgGaS2 and Derivatives: Design, Synthesis, and Optical Properties

College of Electronic and Optical Engineering and College of Flexible Electronics (Future Technology), Nanjing University of Posts and Telecommunications, Nanjing 210023, China
*
Author to whom correspondence should be addressed.
Submission received: 30 December 2024 / Revised: 17 January 2025 / Accepted: 18 January 2025 / Published: 20 January 2025
(This article belongs to the Special Issue Nonlinear Optics and Ultrafast Lasers in Nanosystems)

Abstract

:
Silver gallium sulfide (AgGaS2) is a ternary A(I)B(III)X(VI)2-type semiconductor featuring a direct bandgap and high chemical stability. Structurally resembling diamond, AgGaS2 has gained considerable attention as a highly promising material for nonlinear optical applications such as second harmonic generation and optical parametric oscillation. In attempts to expand the research scope, on the one hand, AgGaS2-derived bulk materials with similar diamond-like configurations have been investigated for the enhancement of nonlinear optics performance, especially the improvement of laser-induced damage thresholds and/or nonlinear coefficients; on the other hand, nanoscale AgGaS2 and its derivatives have been synthesized with sizes as low as the exciton Bohr radius for the realization of potential applications in the fields of optoelectronics and lighting. This review article focuses on recent advancements and future opportunities in the design of both bulk and nanocrystalline AgGaS2 and its derivatives, covering structural, electronic, and chemical aspects. By delving into the properties of AgGaS2 in bulk and nanocrystalline states, this review aims to deepen the understanding of chalcopyrite materials and maximize their utilization in photon conversion and beyond.

1. Introduction

In recent years, semiconductors that adopt a chalcopyrite crystal structure and are characterized by the formula A(I)B(III)X(VI)2 (A = Cu, Ag; B = Ga, In; X = S, Se or Te) have garnered significant attention due to their attractive applications in nonlinear optical devices, photon detectors, light-emitting diodes, and solar cells. These materials possess several advantages, including low toxicity, high structural adaptability, and suitable bandgaps [1,2,3,4,5,6,7,8,9]. Among the diverse chalcopyrite compounds, AgGaS2 stands out, with its transparency window spanning from 0.45 to 13 µm and a substantial bulk bandgap of approximately 2.7 eV. Consequently, AgGaS2 exhibits exceptional optical properties, both linear and nonlinear, making it well suited for applications such as photon upconversion, second harmonic generation, and optical parametric oscillation. These capabilities allow AgGaS2 to promote photon conversion across a wide spectral range while offering solid-state convenience, reliability, and the ability to handle high-power pulses.
Chalcopyrite AgGaS2 belongs to the I-42d space group, in which each Ag+ or Ga3+ is coordinated tetrahedrally by four S2− and, correspondingly, each S2− is also tetrahedrally coordinated by two Ag+ and two Ga3+ cations with different bond lengths (dAg-SdGa-S) to form a tetragonal structure (Figure 1a). Interestingly, the unit cell of titled chalcopyrite AgGaS2 can be constructed from two zincblende unit cells aligned in the c-direction. The inequality between dAg-S and dGa-S is a direct result of different interaction strengths between Ag+/S2− and Ga3+/S2− pairs. From the perspective of crystallography, the atomic positions of an ABX2-type chalcopyrite unit cell can be defined as A (0, 0, 0), (0, 1/2, 1/2); B (1/2, 1/2, 0); (1/2, 0, 1/4), X (u, 1/4, 1/8), (−u, 3/4, 1/8), (3/4, u, 7/8), (1/4, −u, 7/8). Here, u is a parameter describing the displacement of the X2− anion from the center of the [A2B2X] tetrahedron (Figure 1b), which can be defined using the following equation [10]:
u = 1 2 c 2 32 a 2 1 16
where a and c are unit cell parameters for the chalcopyrite ABX2, and η = c/2a is defined as the geometrical parameter. When considering a standard zincblende structure, it is easy to conclude that u equals 1/4. Thus (u − 1/4) is regarded as anion displacement, directly related to the near-neighbor cation–anion distances, as shown in the following equations:
u 1 4 = d A X 2 d B X 2 a 2
d A X = a u 2 + 1 + η 2 16 and   d B X = a u 1 2 2 + 1 + η 2 16
Here, η = c/2a can be used to describe the compression of the chalcopyrite unit cell (η = 1 for a zincblende unit cell). Experimental measurements have determined the Ag–S and Ga–S distances in AgGaS2 to be 2.556 ± 0.001 and 2.276 ± 0.001 Å, respectively [11]. The values of η (ranging from 0.89 to 0.93) and u (ranging from 0.28 to 0.30) have been documented by Laksari’s and Piasecki’s research groups [10,12]. From the perspective of crystallography, the non-centrosymmetric nature of the tetragonal structure (η ≠ 1) is crucial for AgGaS2’s exceptional performance in the field of nonlinear optics (NLO), as it enables birefringence and phase matching. Notably, AgGaS2 also exhibits superior structural rigidity due to the significant difference in ionic radii between Ag+ (1.0 Å) and Ga3+ (0.47 Å). This prevents cation intermixing or substitution, which is commonly observed in other chalcopyrite crystals like CuInS2, where the ionic radii are similar (0.74 Å for Cu+ and 0.76 Å for In3+) [13]. By employing a generalized gradient approximation method to calculate the electronic band structure of AgGaS2, Piasecki and colleagues found that the crystal displayed a nonzero direct energy gap at the Γ point (Figure 1c) [12]. The calculated bandgap (Eg) using the GGA functional was determined to be 1.036 eV, slightly underestimating the experimental value of 2.65 eV.
Figure 1. (a) Crystal structure of AgGaS2 and the [GaS4], [AgS4], and [Ag2Ga2S] tetrahedra. (b) S2− displacement in chalcopyrite AgGaS2. (c) Energy levels of chalcopyrite AgGaS2 exhibiting a nonzero direct energy gap at the Γ point. Reprinted with permission [12]. Copyright 2022, Wiley-VCH GmbH.
Figure 1. (a) Crystal structure of AgGaS2 and the [GaS4], [AgS4], and [Ag2Ga2S] tetrahedra. (b) S2− displacement in chalcopyrite AgGaS2. (c) Energy levels of chalcopyrite AgGaS2 exhibiting a nonzero direct energy gap at the Γ point. Reprinted with permission [12]. Copyright 2022, Wiley-VCH GmbH.
Nanomaterials 15 00147 g001
The pioneering study of AgGaS2 bulk crystals can be traced back to the 1970s. The fundamental investigation of its NLO properties has stimulated intensive work on AgGaS2-related NLO crystals, especially in the infrared wavelength. In contrast, the development of AgGaS2 nanocrystals has faced significant challenges, primarily due to the complex wet-chemical synthesis required to obtain high-quality nanocrystals. Given the growing knowledge of reaction dynamics, the recent growth in research activity on wet chemistry has enabled the preparation of AgGaS2 nanocrystals with decent mono-dispersion and anticipated performance. The low toxicity, facile preparation, and flexible optical tunability of AgGaS2 nanocrystals make them promising for optoelectrical and lighting applications. In addition to the good optoelectrical properties, AgGaS2 has also inspired a series of derivative materials due to its intriguing NLO parameters. Therefore, the need for an in-depth study of AgGaS2 is self-evident, not only to provide an important scientific basis for future material optimization but also to help consolidate its position in existing applications.
Different from the currently existing reviews that mainly focus on the nonlinear properties of chalcogenides, we aim to provide a comprehensive overview of the current progress and future prospects in the design of AgGaS2 and its derivatives in this review. We focus on recent advancements in the understanding of their structural, electronic, and chemical properties. The review begins with an introduction to the synthesis of bulk AgGaS2 crystals in Section 2, which serves as the foundation for their utilization as NLO materials. Section 3 examines the NLO performance of AgGaS2 and its derivatives, presenting the guiding principles for the exploration of high-quality NLO materials. In Section 4, the chemical synthesis of AgGaS2 nanocrystals is discussed, highlighting the selection of appropriate reaction precursors and conditions. Finally, Section 5 summarizes the techniques employed to optically tune AgGaS2 nanocrystals through elemental alloying and core–shell engineering.

2. Bulk Crystal Synthesis

AgGaS2 is a visually striking crystal with a transparent yellow color. It exhibits exceptionally high reflection optical activity, surpassing many other known materials [14]. With a melting point of ~996 °C (Figure 2a) [15], AgGaS2 single crystals are typically synthesized through melt growth using the vertical Bridgman or Czochralski method. From the pseudo-binary phase diagram, it is seen that AgGaS2 has good thermal stability at relatively high temperatures. It lies in a single-phase region at temperatures below 750 °C, above which it departs from the vertical line in the direction of Ga2S3. It is evidenced that single crystals grown from polycrystalline materials with near-stoichiometric compositions may contain excess Ga2S3. In an early study conducted by Route et al. in 1974, AgGaS2 single crystals with a diameter of 1.1 cm and length of 5 cm were successfully grown [16]. However, these crystals often suffered from issues such as the presence of precipitates, twinning, and even cracking, despite achieving optical clarity. The strong optical scattering caused by these imperfections hinders the practical application of low-quality crystals. Therefore, the primary challenge in the development of AgGaS2-based devices lies in the difficulty of growing sufficiently large crystals with high optical transparency [17,18].

2.1. Negative Thermal Expansion of AgGaS2 Crystals

Abnormal thermal expansion behavior during the crystal growth process is one of the main factors contributing to the difficulty of growing high-quality AgGaS2 crystals. It has been observed that AgGaS2 exhibits negative thermal expansion along the c axis when it cools from its melting point to room temperature, which often leads to the cracking of crystal ingots [19,20]. The exact thermal expansion coefficients αa and αc were recently determined by Huang and coworkers with a thermal dilatometer in the temperature range of 298–850 K [18]. The authors showed that αc was −1.14 × 10−5 K−1 at 600 K, and the numerical value of αa was larger than αc in the whole temperature range (αa > |αc|) (Figure 2b). This phenomenon can be attributed to the Ag/Ga vibrations of low-frequency transverse modes as well as the bond length change for dAg-S and dGa-S because the Rietveld refinement of in situ XRD patterns showed an increase in dAg-S and a decrease in dGa-S over the whole temperature range. As negative thermal expansion is an intrinsic nature of AgGaS2 crystals, slowing down the cooling steps along with the use of directed seed AgGaS2 may contribute to the success of growing high-quality AgGaS2 crystals. Current technology has allowed for AgGaS2 growth by using seed crystals oriented with the c axis parallel to the axis of the ampoule in which the crystal grows to avoid the cracking of the ampoule.

2.2. Second-Phase Precipitation of AgGaS2 Crystals

The other reason for the difficulty in growing high-quality AgGaS2 crystals is the occurrence of second-phase precipitation during cooling. According to the phase diagram, AgGaS2 crystallizes and grows in the two-phase region that is deviated from the stoichiometric line (Figure 2a) [15]. As a part of AgGaS2 will separate into Ga2S3 and Ag2S, AgGaS2 single crystals grown from a near-stoichiometric composition may contain excess Ga2S3 to form the intermediate phase Ag2Ga20S31 (Ga2S3 dissolving in solid AgGaS2), leading to the segregation of a thin, black Ag2S layer that deposits on the surface of AgGaS2 ingots (Figure 2c). The excessive precipitation of Ag2S further generates a fine-textured second phase containing Ga2S3 impurities due to supersaturation upon cooling at around 790 °C, which deteriorates the transparency [17]. Fortunately, the black Ag2S layer can be removed by mechanical polishing. Advances in crystal growth with suitable post-treatment have made large AgGaS2 single crystals with improved transparency available as well. For example, the inner Ga3S2 precipitates can be cleared by annealing at a temperature slightly below melting temperature in evacuated quartz chambers containing Ag2S. Moreover, the inner Ga3S2 precipitates can also be removed through a combination of static and dynamic vacuum annealing to balance sublimation and diffusion, as reported by Petit and coworkers [17]. Despite only a trace amount of Ga2S3 precipitates being removed during the dynamic stage (lasting for a few seconds) at 830 °C, the subsequent diffusion of excessive Ga and S species can slowly take place from the center of the bulk to the surface during the static stage (last for a few days). After several cycles, the concentration of Ga2S3 impurities could be decreased to an acceptable level. The successful preparation of large-size AgGaS2 crystals with few defects enables fundamental research on their NLO performance, and now AgGaS2 crystals are available on the commercial market [21,22].

3. AgGaS2 and Derivatives for Nonlinear Optics

NLO materials have been of great interest in recent decades because of their spectacular applications such as space communications, infrared countermeasures, medical diagnosis, and remote gas spectroscopy [23,24,25,26,27,28,29,30]. Governing SHG and phase-matching mechanisms are core concepts in modern NLO. The realization of phase matching is the key to ensure the efficiency of SHG, and its goal is to keep the phase of the incident light and its corresponding second harmonics consistent when propagating through the medium to reduce energy loss. Crystals belonging to the chalcopyrite symmetry class were proposed as possible infrared tunable optical parametric oscillation (OPO) sources as early as 1971 [31]. AgGaS2 crystals are transparent in the wavelength region of 0.45–13 µm with a nonlinear coefficient d36 of 11–13 pm/V and a birefringence Δn of 0.053 at 10.6 µm [32]. Note that such an adequate birefringence enables phase matching for both type-I (two low-frequency waves with the same polarization) and type-II (two low-frequency waves with orthogonal polarization) interactions, which are described by the following equations:
f o r   t y p e   I :   d e f f = d 36 s i n θ s i n 2 φ
f o r   t y p e   I I :   d e f f = d 36 s i n 2 θ s i n 2 φ
where deff is the effective nonlinear coefficient, d36 is the nonlinear tensor component of AgGaS2, and θ and φ are the polar and azimuthal angles, respectively. Taking advantage of NLO properties, AgGaS2 crystals have been developed for frequency doubling, frequency mixing, and as parametric oscillators so that continuously tunable radiation from 1.0 to 12 µm can be produced under suitable pumping laser excitation. Notably, AgGaS2 is one of the few crystals that can be pumped by commercially available 1.06 µm lasers to achieve phase-matched downconversion into the λ > 5 µm region.
The groundbreaking utilization of AgGaS2 in OPO was first reported by Byer and colleagues, marking the first record of OPO in this class of chalcopyrite semiconductors [33]. The authors fabricated AgGaS2 crystals measuring 2 cm in length and 1.0 × 0.5 cm2 in cross-section, cut at an angle of 50.5° on the optical axis in the [110] direction. Pumped by an electro-optic Q-switched YAG: Nd laser (20 ns, 10 Hz), AgGaS2 OPO could achieve tunable emission in the wavelength range of 1.4 to 4 µm with type-I interactions. Although the operating wavelength of the AgGaS2 OPO could extend to the crystal’s transparency limit (~12 µm), the parametric oscillation threshold could not surpass λ > 4 µm due to significant damage occurring at the crystal surface. Subsequently, Vodopyanov and colleagues achieved the experimental realization of AgGaS2-based OPO with a wide mid-infrared tuning range [34]. The authors utilized AgGaS2 crystals measuring 2 cm in length and 0.7 × 1.0 cm2 in cross-sectional area. Narrow OPO linewidths were achieved, and effective second-order nonlinearity was maximized as well by choosing type-II phase matching, for which θ~45° yields deff = 12 pm/V, 25–30% higher than in the case of type-I (θ~52°). As a result, the singly resonant type-II AgGaS2 OPO pumped by a nanosecond YAG: Nd laser could yield an idler wave continuously tunable from 3.9 to 11.3 µm with a linewidth of 1 cm−1 and an OPO threshold of 0.03 J/cm2.
Although AgGaS2 crystal possesses a wide transparency window and a large second harmonic generation (SHG) effect, its laser-induced damage threshold (LIDT) is rather small (that is, ~10 MW/cm2@1.06 µm) according to Vodopyanov’s report [34], owing to its relatively narrow bandgap (2.7 eV). Other commercially available infrared NLO materials such AgGaSe2 and ZnGeP2 also exhibit large SHG responses but still encounter the issue of low LIDT or two-photon absorption (TPA) [35,36]. The TPA effect of these materials will cause energy loss under a high power input, thus reducing the conversion efficiency. Moreover, Ag-containing NLO materials like AgGaS2 additionally suffer from the photodarkening effect of elemental Ag, which dramatically shortens the crystal longevity. These drawbacks to some extent hinder their practical applications for long-term use. In this regard, it is desirable to explore novel alternative NLO materials that exhibit better NLO performance in the infrared spectral region [37,38,39,40]. In the literature, it is well recognized that the optical properties of an ideal infrared NLO material should meet the following criteria: (i) the infrared transparency window should be broad enough to at least cover two important atmospheric transparent windows (i.e., 3–5 µm and 8–12 µm); (ii) the SHG coefficient dij should be large, at least larger than 10 × KDP (KH2PO4, with d36 = 0.39 pm/V); (iii) the LIDT should be higher than that of AgGaS2, and the corresponding bandgap should be larger than ~3.0 eV; (iv) the birefringence Δn should fall in a suitable range (i.e., 0.03–0.10) to achieve the NLO phase-matching condition; and (v) the crystal should have a good growth habit and chemical stability. Taking chalcopyrite AgGaS2 as a model, the alternatives are mainly focused on its derivatives possessing a similar diamond-like structure (Figure 3) [41].
In a classic diamond-like ZnS material, the Zn–S bond lengths within the [ZnS4] group have a narrow range of 2.339–2.343 Å, with only a slight difference of 0.004 Å. The S–Zn–S angles are located in the range of 109.65–109.68°, which has only a very small deviation from the standard tetrahedral structure (109.5°). As a result, binary diamond-like materials such as ZnS are not suitable for infrared SHG applications due to their limited structural anisotropy, resulting in small birefringence [37]. As an alternative, diamond-like ternary or quaternary compounds derived from AgGaS2 with suitable phase-matching properties are more promising [42,43]. These compounds exhibit relatively weak refractive index dispersion in the long wavelength region, allowing for a relatively small birefringence (Δn ~ 0.03–0.07) to satisfy the phase-matching condition in the UV-to-infrared spectral range. Excessive birefringence (Δn > 0.10) can lead to destructive optical frequency conversion due to large walk-off and self-focus effects, which impede the generation of high-power coherent light [44,45]. Various sulfide units, such as [PS4], [SiS4], [GeS4], [InS4], [CdS4], [ZnS4], and [LiS4] can be substituted for the fundamental tetrahedral building blocks [AgS4] and [GaS4] in the diamond-like 3D framework. These tetrahedra can be interconnected and aligned, enabling the additive superposition of the microscopic second-order susceptibility tensors. It is worth noting that finding an ideal NLO material with a wide energy gap (Eg > 3.0 eV) and a large SHG efficiency (greater than that of AgGaS2) is highly desirable but remains a significant challenge due to the inherent trade-off between the SHG coefficient and bandgap value. Increasing the bandgap excessively will lead to a decrease in the SHG coefficient. To this end, the balanced SHG coefficient and bandgap value of AgGaS2 provide it with good application prospects. In the pursuit of efficient and reliable NLO materials, AgGaS2 is obviously an important reference. Further advancements in performance and quality focus on modifying the bandgaps or SHG coefficients, as summarized in Table 1.

3.1. Bandgap Modification

The narrow bandgap of AgGaS2 is attributed to the presence of the 3d orbitals of Ag at the valence band maximum [37,58,59,60,61]. To modify the bandgap and improve the LIDT, the incorporation of an electropositive element, such as an alkaline metal (e.g., Li and Na) or alkaline earth metal (e.g., Ba), into the Ag sites is considered. These elements do not exhibit d-d or f-f electron transitions [62]. For instance, replacing [AgS4] with [LiS4] results in the formation of LiGaS2, which possesses a significantly larger bandgap of 3.7 eV and exhibits a notably higher LIDT (11 times that of AgGaS2) [51]. However, when Li+ is substituted for Ag+ in AgGaS2, the space group of AgGaS2 collapses into a lower Pna21 orthorhombic symmetry. As a result, the NLO coefficient of LiGaS2 is significantly reduced (d36 = 5.8 pm/V), and its SHG performance is inferior to that of AgGaS2. Considering that Li+ and Ag+ share many chemical similarities, such as the same valence state and four-fold tetrahedral coordination in a compound, some researchers have suggested that Li+ might be dissolved in AgGaS2 to form a solid solution, while the space symmetry could be preserved. This idea was proposed by Wu and coworkers, who examined a series of LixAg1-xGaS2 crystals in 2019 [53]. The authors found that for LixAg1-xGaS2, Li+ substitution did not change the crystal structure when x was tuned from 0 to 0.6. Correspondingly, the bandgaps can be fine-tuned from 2.58 (x = 0, AgGaS2) to 3.40 eV (x = 0.6, Li0.6Ag0.4GaS2) (Figure 4a). The large bandgap of Li0.6Ag0.4GaS2 enables an 8.6-fold enhancement in the LIDT at 1.06 µm compared to that of AgGaS2. This improvement may be attributed to the reduced difference in cationic radii between Ag+ (1.0 Å) and Ga3+ (0.47 Å) through Li+ (0.59 Å) substitution. In mixed-cation crystals like Li0.6Ag0.4GaS2, dislocations can be modified, shifting the structure back toward the ideal diamond-type structure and allowing for better alignment of the [GaS4] units along the chalcopyrite direction.
In addition to the Li+ substitution, Na+ or Ba2+ substitution is also feasible. It should be noted that the introduction of the Na+ cation will break the 3D [GaS4] framework into 2D layers [63,64]. In NaGaS2 crystals, Ga atoms are four-coordinated with S atoms to form [GaS4] tetrahedra, and every four tetrahedra gather to form a super-tetrahedral [Ga4S10] cluster through corner sharing (Figure 4b). The unique packing configuration results in a large bandgap of 3.9 eV and a high birefringence of ~0.09 [48]. Meanwhile, the blue shift of the UV cutoff edge to 307 nm is observed, suggesting promise as a functional material working in the visual-to-near-IR region. However, the authors have not characterized the SHG coefficient dij of the NaGaS2 crystals. By inspecting the symmetry of NaGaS2, there are reasons to believe that the SHG performance cannot surpass that of AgGaS2. The introduction of alkaline earth elements such as Ba into the diamond-like sulfide framework can also modify the bandgap [65]. For example, Dong and coworkers found that Ba2Ga8SiS16 and Ba2Ga8GeS16 adopted a 3D framework structure exhibiting the alternate stacking of two distinct tetrahedral layers (Figure 4c) [66]. The samples adopted the same P63mc space group and had large bandgaps (3.4 eV for Ba2Ga8SiS16 and 3.0 eV for Ba2Ga8GeS16) and a broad transparency window (0.42–20 µm). Particularly, Ba2Ga8GeS16 exhibited an LIDT ~22 times higher than that of AgGaS2 owing to the large bandgap. Ba2Ga8GeS16 also exhibited strong SHG signals that were comparable to those of the benchmark AgGaS2 at a laser irradiation of 1.95 µm. The strong nonlinear effect was presumably related to the synergic effect of the alternate stacking of the mixed [(Ga/Ge)S4] and the pure [GaS4] tetrahedral layers along the c axis, the alignment of these two types of tetrahedra in the same direction, and the slightly off-center feature of alkaline earth cations in the interstitial sites.
Apart from alkaline and alkaline earth substitution, the introduction of P5+ into chalcogenide polyhedra has also been reported to enlarge bandgaps [67]. Metal thiophosphate can intrinsically satisfy the crucial criteria for a promising infrared NLO crystal due to the strong covalent P–S bonds [68,69]. Moreover, thiophosphate exhibits diverse P–S anionic groups such as [PS4] and [P2S6] [70], which can further connect with other NLO-active units to construct varied crystal architectures, giving rise to more possibilities for designing NLO crystals. It should be mentioned that thiophosphates usually possess a relatively low melting point, and this feature certainly favors large crystal growth. Guided by this idea, Yao and coworkers designed a novel thiophosphate CuZnPS4 that showed a sharply enlarged energy gap (3.0 eV) and reinforced the SHG response (d14 = 15.9 pm/V) (Figure 4d) [67]. The lengths of the Cu/Zn–S and P–S bonds were 2.3227 and 2.0426 Å, which were shorter than those of Ag–S (2.605 Å) and Ga–S (2.235 Å) bonds in AgGaS2, implying stronger covalent interactions in the CuZnPS4. With the assistance of density functional theory calculation, the authors indicated that the highest occupied atomic orbitals (HOMO) of CuZnPS4 were dominated by earring-shaped S 3p electrons, while the lowest unoccupied atomic orbitals (LUMO) primarily comprised S 3p and P 3p electrons. Owing to the occurrence of nonbonding states, the electron density in the adjacency of the forbidden gap of CuZnPS4 was significantly increased in comparison to that of AgGaS2.
The introduction of an alkaline metal (Li+ and Na+), alkaline earth metal (Ba2+), or P5+ modifies the bandgap by changing the crystal structure and the local electronic environment. However, the introduction of additional elements may lead to the distortion of the crystal structure, especially at high concentrations. This leads to a decline in optical properties and a weakening of material stability under fluctuant environmental factors such as humidity and temperature.

3.2. SHG Modification

The SHG response is largely related to the coordination and polarization of the polyhedron. Compounds containing transition-metal cations (e.g., Cu+, Zn2+, Cd2+, and Hg2+) with an (n-1)d10ns0 configuration or main-group metal cations (e.g., Pb2+, Sn2+, Bi3+, and Te4+) with an ns2 configuration are potential candidates for polar crystals. It is known that the former elements have a polar displacement of d10 cation centers, while the latter ones undergo the second-order Jahn–Teller (SOJT) effect caused by lone-pair electrons [71]. Among these tetrahedra, Hg2+-containing diamond-like sulfides have been widely studied because Hg is the heaviest member in the d10 group, and Hg–S bonds are highly polarizable, which is believed to significantly increase the SHG response and birefringence. In addition, [HgS4] tetrahedra exhibit a larger distortion compared to other tetrahedra like [ZnS4] and [CdS4] [72]. As a typical example, Yao and coworkers reported the replacement of [GaS4] units by highly distorted [HgS4] and [PS4] units for the preparation of the new NLO material AgHgPS4 with a bandgap of 2.63 eV (Figure 5a,b) [56]. The authors calculated the contributions of [AgS4] (−10.64 pm/V), [HgS4] (−30.77 pm/V), and [PS4] (−11.54 pm/V) units and concluded that the [HgS4] units contributed most to the phase-matching SHG response (d11 = −31.08 pm/V). Moreover, they demonstrated that this material system could be further extended to a uniform structure of AHgPS4 (A = Li and Cu), demonstrating the great advantages of the combination of cation vacancy defects and highly distorted [HgS4] tetrahedra. In a parallel contribution, Wu and coworkers reported designed and synthesized Li2HgMSe4 (M = Si, Ge, and Sn) NLO materials by concurrently replacing the cation Ag+ and the [GaSe4] unit with the alkali metal Li+ and [HgSe4]/[MSe4] to optimize crystal structures and performances (Figure 5c,d) [57]. Optical characterization showed that Li2HgMSe4 exhibited bandgaps of 1.2–1.7 eV and strong powder SHG responses (3.6–6.0 × AgGaS2) with the essential phase-matching behavior (Δn = 0.042 − 0.172). Particularly, Li2HgSnSe4 exhibited the largest SHG response (d33 = −104.3 pm/V) among the known Hg-based chalcogenides. This research highlights the practicability of the functional group co-substitution strategy and shows that Hg-based chalcogenides can be a promising system for the future exploration of large SHG crystals.
It is known that the distribution of valence electrons and the spatial arrangement of active groups are two determinants of NLO response. Although the [AgS4] and [GaS4] tetrahedra in AgGaS2 already occur in a favorable way with co-parallel alignment and high density, there is still room for improving NLO performance by fine-tuning the valence electron distribution through doping or by designing tetrahedral positions through multiple doping.
According to the valence electron concentration (VEC) theory, the VEC can be defined using the following equation [67]:
V E C = m e A + n e B + m + n +
where eA and eB are the valence electron numbers of compositional elements A and B in the [AB4] tetrahedral unit; m and n represent the atomic numbers of A and B, respectively. Guided by this principle, a normal diamond-like structure such as that of AgGaS2 adopts a VEC number of 4. With the introduction of an element with high electrovalence such as Ge4+ in a normal diamond-like compound, the VEC number will obviously increase owing to the occurrence of cation vacancies. For example, Hg2GeSe4 shows a large VEC number of 4.57 [73]. By first-principle calculation, Lin and coworkers found that Hg2GeSe4 has a bandgap of 2.8 eV, a moderate birefringence value of 0.096, and a large SHG coefficient of d36 = 38.22 pm/V [71].
The idea of the spatial arrangement of active groups can be realized by combining two or more different typical NLO-active motifs, such as [GaS4], [BaS4], [InS4], [GeS4], and [SiS4] tetrahedra, forming super-polyhedral clusters in a single chalcogenide compound to modulate SHG efficiency and LIDT simultaneously [52,74,75]. One can imagine that multiple polyhedra in the structure will lead to large SHG responses. Guo and coworkers presented two infrared NLO materials, Na2Ga2GeS6 and Na2Ga2SnS6, by mixing the different typical NLO-active motifs [GaS4] and [GeS4]/[SnS4] into the alkali metal-containing system (Figure 6a) [55]. Optical characterization showed that Na2Ga2GeS6 and Na2Ga2SnS6 had optical bandgaps of 3.1 and 2.74 eV and deff values of 11.18 and 13.11 pm/V. In another work guided by the strategy of co-substitution, Zhou and coworkers reported the preparation and optical characterization of Li2ZnSiS4 crystals [49]. The structure of Li2ZnSiS4 can be considered a replacement of the Ag and [GaS4] units by alkali metal Li [ZnS4]/[SiS4] units (Figure 6b) [76]. Theoretical calculations and optical characterizations indicated that Li2ZnSiS4 showed a large bandgap of 3.9 eV, a broad transparency window range of 3–25 µm, a high LIDT (10 × AGS), and a large phase-matching SHG response of d33 = 18.89 pm/V at 2.09 μm. This co-substitution idea was further extended to the salt-inclusion chalcogenides [RaXb][MS4] (R = alkaline or alkaline earth elements; X = halide elements; and M = main-group or/and transition elements), where tetrahedral [MS4] works as an asymmetric building unit, and ionic guest [RaXb] provides filler ions to stabilize the structures [77,78,79]. For example, Wu and coworkers designed the salt-inclusion chalcogenide [Ba4Cl2][ZnGa4S10] in accordance with the AgGaS2 template (Figure 6c) [50]. Structural characterization showed that [Ba4Cl2][ZnGa4S10] was constructed from corner-sharing super-tetrahedral [Ga4S10] clusters and [ZnS4] tetrahedra to form a three-dimensional open [ZnGa4S10] diamond-like framework. Ba2+ and Cl ions interpenetrated the framework to support charge balance. Further optical characterization indicated that the [Ba4Cl2][ZnGa4S10] crystal possessed a wide bandgap of ~3.85 eV, a large SHG coefficient of d14 = 14.9 pm/V, and an ultrahigh LIDT of 51 × AGS, demonstrating its promise as an infrared NLO candidate.
It should be emphasized that all diamond-like ternary or quaternary materials are SHG active because of their inherently non-centrosymmetric structures. However, when the overall performances are compared, it seems that AgGaS2 is still one of the most useful NLO materials since it possesses balanced NLO properties [40,80]. Therefore, many studies are still focused on the classic AgGaS2. For example, Cheatum and coworkers presented a method for increasing the power of mid-infrared laser pulses generated by AgGaS2 difference frequency generation, amplifying mid-infrared light twofold [28]. Indeed, most of the other known ternary and quaternary diamond-like structure materials exhibit either smaller bandgaps or weaker SHG effects [81]. One particular material, that is, Li2ZnSiS4, needs further systematic study since it presents a superior balance between the bandgap (>3.0 eV) and the SHG effect (>AgGaS2), as mentioned in the above discussion [41]. This result implies that although it seems challenging to find an optimized infrared NLO material, it is possible to design a target material by using AgGaS2 as a model for quaternary diamond-like metal sulfides.

4. Nanocrystal Synthesis

The traditional solid-state reaction for synthesizing AgGaS2 involves high processing temperatures (800–1000 °C), long reaction times, strict sintering processes, and specialized equipment. In contrast, the wet-chemical synthesis of AgGaS2 nanocrystals is more straightforward and time-saving. By reducing the crystal size to the excitonic Bohr radius (3.3 nm for AgGaS2) [82,83,84,85,86], the restricted electron movement results in wideband light absorption and bright light emission, known as the quantum confinement effect [87,88]. Moreover, the large surface-to-volume ratio of nanocrystals allows for flexible surface modification, leading to extended functionalities. While there has been extensive research on chalcopyrite semiconductor nanocrystals such as CuInS2 [89], CuGaS2 [90], and CuInSe2 [91], AgGaS2 nanocrystals are not as well explored, mainly due to the challenges associated with incorporating Ag.

4.1. Solvothermal Synthesis

Early attempts to synthesize nanocrystalline AgGaS2 were made by Qian and coworkers in the late 1990s [92]. Through a standard solvothermal synthetic route, the authors prepared tetragonal AgGaS2 nanocrystals 5–7 nm in diameter by using AgCl, Ga, and S as the reactants at 180–230 °C. In a modified reaction, they used AgCl, GaCl3, and thiourea as reactants [93]. Interestingly, the introduction of thiourea interacted with AgCl to form a water-soluble Ag–thiourea complex in a homogenous solution, which was helpful to form nanocrystalline AgGaS2 (~5 nm in diameter). Consequently, spectral results showed that the photoluminescence spectrum consisted of one broad emission feature at 446 nm, which was blue shifted at about 50 nm compared with that of single-crystal AgGaS2, indicating the quantum confinement effect. However, the AgGaS2 nanocrystals prepared via the hydrothermal route at that time were inclined to form assembled microparticles, attributed to the lack of proper surface ligands. Through a solvothermal synthesis, AgGaS2-based nanohybrids can be easily formed. In a recent study, Wang and colleagues found that the attachment of AgGaS2 nanosheets to the edges of carbon nitride resulted in the formation of a type-II heterostructure, exhibiting an enhanced photocatalytic property [85]. The following contributions mainly focus on the synthesis of monodispersed AgGaS2 nanocrystals though bench-top methods employing heat-up and hot-injection techniques (Table 2).

4.2. Heat-Up Synthesis

The successful growth of AgGaS2 nanocrystals through a heat-up procedure typically involves fast nucleation and subsequent crystal growth, as described by the classic LaMer model [106]. Compared to the classic binary quantum dots, the formation of uniform AgGaS2 nanocrystals in the Ag–Ga–S ternary system is more complicated because the reactivity of double cations should be balanced in addition to temporally separating the nucleation and growth processes. To restrain dual nucleation processes, the reaction activities of two metal precursors should not differ significantly from each other, which is identified as the key to deliberately tuning the electronic band feature of the nanocrystals.
A previous study reported that when single-source AgGa(S2CN(C2H5)2)4 was used as the precursor and oleylamine was used as the solvent, AgGaS2 nanocrystals could be obtained from a reaction at 180–240 °C for 10 min [97]. Later, Han and coworkers reported the synthesis of AgGaS2 nanocrystals using AgS2CN(C2H5)2 and Ga(S2CN(C2H5)2)3 complexes separately in conjunction with proper coordination solvents [94]. Structural characterization showed that pure-phase AgGaS2 nanocrystals were obtained when long-chain alkanethiols (dodecanethiol and hexadecanethiol) or alkylamines (oleylamine and hexadecylamine) were employed (Figure 7). It should be noted that in this case, the final products were in an orthorhombic phase rather than a common tetragonal phase. When oleic acid was employed as solvent, AgGaS2 nanocrystals were not successfully prepared, and instead, only a mixture of Ag and Ag2S was produced; when a tertiary alkylamine such as trioctylamine was utilized, a mixture of tetragonal AgGaS2 and Ag9GaS6 was produced; and if trioctylphosphine was used, only Ag crystals were yielded without any Ga-associated byproducts. Moreover, although the preparation of AgGaS2 nanocrystals was feasible using this method, the crystal sizes were excessively large (>15 nm); therefore, they were beyond the scope of the quantum confinement, and thus little photoluminescence was recorded.
In addition to the synthesis of AgGaS2 nanocrystals with a hydrophobic nature, the direct synthesis of nanocrystals with a hydrophilic surface is favored for biomedicine-related applications. This idea was realized recently in a demonstration reported by Azhniuk and coworkers [96]. The colloidal synthesis of AgGaS2 nanocrystals was carried out in a coprecipitation reaction between Na2S and a mixture of Ag(I) and Ga(III) thiolate complexes in aqueous glutathione solutions at 93–98 °C. Glutathione acted as not only the S source but also as the coordinating ligand so that small AgGaS2 nanocrystals with a size of ~2 nm could be obtained, and a blue shift of the photoluminescence peak was observed in comparison to that of the reported large nanocrystals.

4.3. Cation-Exchange Synthesis

To control crystal size, synthesizing AgGaS2 nanocrystals through a cation-exchange method was explored by Hughes and coworkers [95]. The formation of AgGaS2 nanocrystals was initiated by the rapid injection of elemental sulfur in dodecanethiol into a reaction vessel containing AgNO3 and Ga(acac)3 in octadecene at 170 °C. The authors observed an immediate change in solvent color from colorless to black and finally to translucent red–brown, indicating the formation of Ag2S at the early stage. This result also implies a potential cation exchange from Ag2S to AgGaS2 during the synthesis, which was proven by the sample characterization of aliquots taken at very early times. A deliberate cation-exchange synthesis method for AgGaS2 was developed by Song and coworkers, who used the as-synthesized AgInS2 nanocrystals as templates (Figure 8) [98]. As a thermal dynamic and kinetic process, nanoscale cation-exchange reactions generally consist of four steps driven by solvation energy and Lewis acid–base interactions: (i) extraction of the surface host cations; (ii) incorporation of guest cations into the nanocrystal surface; (iii) diffusion of the surface guest cations into the interior of nanocrystals; and (iv) diffusion of the inner host cations to the nanocrystals surface. In this regard, the cation exchange rate is in a close relationship with the cation concentration. In the demonstration by Song’s team, at an initial ratio of [Ga]/[In] = 3.73, the In3+ ions were entirely replaced by Ga3+ ions, leading to the formation of AgGaS2 with a stoichiometry of [Ag]/[Ga]/[S] = 1: 0.98: 2.11. Owing to the small ionic radius of Ga3+ (0.47 Å) in comparison with that of Ag+ (1.0 Å) and the strong bond strength of Ag–S, the Ag framework was well preserved during the cation-exchange process. The photoluminescence peak of the resulting nanocrystals could be tuned from 502 to 719 nm by regulating the reaction conditions, with the highest quantum yield of up to 37%.
We understand that for small nanocrystals like quantum dots, the crystal sizes and morphologies play a great role in their optical properties, attributing to the quantum confinement effect. Indeed, the preparation of AgGaS2 nanocrystals, especially the role of surface ligands in stabilizing AgGaS2 nanocrystals, is still an area that requires further exploration and research. Small changes in the reaction conditions (such as pH, temperature, reaction time, etc.) in wet chemistry may affect the properties of the product, which requires strict process control and optimization. Currently, there are limited data available on the synthesis of AgGaS2 nanocrystals, whether through direct synthesis or cation-exchange methods. Due to the potential use of AgGaS2 nanocrystals as blue lighting sources, there is a high demand for the development of wet-chemical synthesis methods. To advance the field, it is important to gain a deeper understanding of the reaction dynamics involved in the formation of AgGaS2 nanocrystals. This can be achieved through in-depth examinations of nanocrystal formation processes, as well as investigations into the composition of the nanocrystals using in situ microscopy techniques. By studying the reaction dynamics and nanocrystal composition, researchers can gain valuable insights into the factors influencing the synthesis process and make improvements to achieve the desired nanocrystal properties. It should be mentioned that with the aid of theoretical and simulation methods and tools, researchers can now guide the synthesis of novel semiconductors using a predictable approach. For example, Filho and coworkers revealed the self-induced formation and stability of InAlN nanorods through ab initio simulations and a density functional theory-based phase-field model [107,108]. It is certainly possible that such a method can be applied to guide the synthesis of AgGaS2 nanocrystals.

5. Optical Tuning of AgGaS2 Nanocrystals for Lighting and Photovoltaic Applications

The limited optical tunability of pristine AgGaS2 nanocrystals can be overcome by employing strategies such as element alloying and core–shell configurations. These approaches allow for precise control of the nanocrystal composition, enabling energy-level engineering and fine-tuning of the emission properties. By introducing alloying elements into AgGaS2 nanocrystals, the energy levels and bandgaps can be modified, leading to tunable emission properties. This alloying process involves incorporating different elements into the crystal lattice, which alter the electronic structure and introduce new energy states. In addition to alloying, the development of core–shell structures in AgGaS2 nanocrystals provides another avenue for emission tuning. The shell can modify the electronic and optical properties of the nanocrystals, leading to wavelength and intensity tuning of the emitted light. By varying the composition and thickness of the shell layer, precise control over the emission properties can be achieved. By leveraging these strategies, it is now possible to engineer energy levels and fine-tune emission properties, expanding the range of potential applications for AgGaS2 nanocrystals.

5.1. Element Alloying for Lighting Applications

Element alloying has been widely adopted to tune the bandgap energy and photoluminescence of semiconducting quantum dots through band mixing, according to Vegard’s law [109]. For AgGaS2, Ga3+, and S2−, sites can be alloyed by In3+ and Se2− ions, which greatly tune their absorption and emission bands. Given the structural similarity of AgGaS2 and AgInS2, Ga3+-site engineering was reported early in 2010 [97]. Kuwabata and coworkers reported the synthesis and emission of AgInxGa1−xS2 (0 < x < 1) nanocrystals via the thermal decomposition of a metal ion–diethyldithiocarbamate complex in oleylamine. Inductively coupled plasma atomic emission spectroscopy (ICP-AES) results confirmed that the ratios of [Ga] to [In] in the resulting nanocrystals were almost equal to those in the precursors, implying the complete thermal decomposition of the precursors. Spectroscopy results showed that the increase in the [Ga]/[In] ratio led to a red shift from 550 to 750 nm in the photoluminescence spectra, attributed to the band mixing of AgInS2 (~1.8 eV) and AgGaS2 (~2.6 eV) (Figure 9a). Notably, the alloyed nanocrystals showed high quantum yields of ~40%, although the pristine AgGaS2 nanocrystals were almost non-luminescent. It should be mentioned that due to the different reaction activities between Ga and In precursors (usually, Ga precursors have a low activity because of their small ionic radius), the cation distribution may not be uniform in the single nanocrystals. For example, it has been observed that AgInS2 nanocrystals could be synthesized at 150 °C, while AgGaS2 nanocrystals formed only at a temperature over 250 °C under the same reaction conditions. In this regard, the balance of cation reactivity is essential in the synthesis of real AgInxGa1−xS2 nanocrystals. Recently, researchers developed a synthetic route using multiple precursors to ensure balance [104]. They used gallium diethyldithiocarbamate (Ga(DDTC)3) as the target precursor, which reacts under moderate conditions to supply both Ga and S. The use of this precursor with silver acetate and indium acetate allowed for the synthesis of AgInxGa1−xS2 nanocrystals at 150 °C.
In addition to In3+ alloying, Zn2+ alloying is also possible, as the ionic diameter of tetragonally coordinated Zn2+ (0.60 Å) is comparable to that of Ga+ (0.47 Å), which thus can be preferentially substituted by Zn2+. For example, Yang and coworkers found that when Zn was alloyed, the emission of AgGaS2: Zn nanocrystals could be tuned from 515 to 450 nm [100]. In another contribution, Tang and coworkers reported the synthesis of narrow-bandwidth blue-emitting AgGaS2: Zn nanocrystals via a facile one-pot method [83]. When the Ag/Zn and Ag/Ga feeding ratios were selected with values of 4:1 and 1:8, respectively, AgGaS2: Zn nanocrystals demonstrated a typical blue emission at 470 nm with narrow full width at half-maximum (FWHM) of 48 nm and a high PLQY of 16.7% (Figure 9b), which was suitable for solution-processed quantum-dot light-emitting diode (LED) applications. It should be noted that due to the difference in chemical reactivities between Ga and Zn, the chemical composition of AgGaS2: Zn nanocrystals is probably still nonuniform. In fact, in one-pot synthesis, Zn is potentially concentrated in the interior of the AgGaS2: Zn rather than forming a shell, since the reactivity of the Zn precursor is a little higher than that of the Ga precursor, which has been evidenced by energy-dispersive X-ray spectroscopy. Despite this inhomogeneity, the role of Zn is still prominent in promoting emission efficiency. Tang and coworkers attributed the emission promotion to the increase in the emission center, induced by this hetero-valent substitution. The addition of an appropriate Zn precursor for one-pot synthesis can produce more Ag+ vacancies, which are responsible for charge carrier recombination. In addition, of course, the introduction of the appropriate amount of Zn can form a layer of ZnS, providing surface passivation for luminescence quenchers.
Apart from the above example, S2−-site alloying with Se2− to form AgGa(S1–xSex)2 is also feasible, and such alloying leads to a red shift in both the absorption and emission spectra [99]. Furthermore, multiple sites can be alloyed simultaneously, thereby greatly expanding the absorption and emission properties [110]. However, due to the unbalanced reactivity, multiple alloying requires more precise control over the uniform growth of alloyed nanocrystals during the entire synthesis procedure, which is promising but yet to be developed.

5.2. Core–Shell Engineering for Lighting and Photovoltaic Applications

The growth of an epitaxial shell on the surface of nanocrystals is a common approach for achieving surface passivation as well as emission tuning, depending on the alignment between core and shell band structures [109,111,112]. The electronic band in a type-I structure is characterized by a straddling band alignment, in which the wider-bandgap shell confines the wave function of the electron–hole pair within the AgGaS2 region. ZnS is the most popular shell material for passivating AgGaS2 nanocrystals due to its large bulk bandgap, proper band alignment, and superior stability. Bai and coworkers developed a one-pot strategy for the synthesis of AgGaS2@ZnS core–shell nanocrystals without prior purification of the as-synthesized AgGaS2 core nanocrystals [101]. Structural characterization was performed to confirm the AgGaS2@ZnS with a designed core–shell configuration. The energy bandgap of the AgGaS2@ZnS core–shell nanocrystals was effectively tunable from 2.98 to 2.83 eV by controlling their non-stoichiometry (Figure 10a). In addition, it could further continuously decrease to 1.90 eV in AgGaxIn1−xS2@ZnS core–shell nanocrystals in conjunction with the In alloying strategy. Benefitting from surface passivation, the resultant core–shell nanocrystals showed largely improved quantum yields of ~30% for LED application.
In addition to the use of ZnS as a shell component, researchers have found that GaSx and CdSeS can also be employed as shells for different purposes. For example, when AgInxGa1−xS2 nanocrystals were coated with GaSx, the original broad defect-site emission peak was drastically and selectively suppressed, leading to an increase in the quantum yield of the band-edge emission peak (Figure 10b) [103]. The optimal quantum yield of AgInxGa1−xS2@GaSx core–shell nanocrystals was 28%, with green band-edge emission at 530 nm and a full width at half-maximum of 181 meV (41 nm). After GaSx coating, the intensity of the defect-site emission was remarkably reduced to less than 15% of the intensity of the corresponding bandgap emission. CdSeS shelling generally leads to inducing a type-II band structure. A red shift of the photoluminescence peak was observed for the AgGaS2@CdSeS core–shell nanocrystals, the emission of which was ascribed to the recombination of delocalized electrons in the CdSeS shell and localized holes within the AgGaS2 core QDs (Figure 10c) [105]. Such AgGaS2@CdSeS core–shell nanocrystals with charge carrier separation features are useful in photovoltaic applications [113]. In another work, Wang and coworkers reported efficient blue LEDs using AgGaS2@GaSx as the light-emitting layer, demonstrating the potential of AgGaS2 quantum dots as a high-quality blue light-emitting material [88].
One may have noticed that during shell growth, a band shift in both absorption and emission spectra in comparison to those of the AgGaS2 core nanocrystals is always observed, implying that the practical outcome of core–shell engineering may result from complex interplays such as etching, cation exchange, and alloying [114]. This phenomenon has also been widely reported in other ternary chalcopyrites like CuInS2 [115]. In addition, either for lighting or for photovoltaic applications, there is a trade-off between shell thickness and performance. Current demand requires core–shell nanocrystals with sufficient shell thickness but suppressed alloying. In this regard, the development of low-temperature deposition techniques may be a potential option. In addition, stoichiometry control is another manner for optical tuning in other ternary systems such as CuInS2 [116,117]; however, this technique has not been well studied for AgGaS2. There are few data concerning the role of the Ag/Ga ratio in optical tuning; 1:8 is adopted in some previous reports [83,100]. Due to the fact that a higher degree of Ag deficiency in AgGaS2 can lower its valence band maximum, since it is composed of the hybrid orbitals Ag 4d and S 3p, the study of stoichiometry control, especially the Ag/Ga ratio, deserves special attention in the Ag–Ga–S system in future work.

6. Conclusions

This review documents the development of bulk and nanocrystalline AgGaS2 as well as their derivatives from the perspective of synthesis, NLO response, and optical tuning. With significant advancements made over the past decade, the study of AgGaS2 has stimulated the further exploitation of a series of diamond-like AgGaS2 derivatives with larger bandgaps, higher LIDTs, or stronger SHG coefficients, expanding the toolbox of infrared NLO materials. In addition, AgGaS2 and derived nanocrystals with suitable sizes and good mono-dispersion can be obtained through direct synthetic control or post-synthesis cation exchange. Through band intermixing, alloyed nanocrystals also provide more possibilities for tuning their absorption and emission features. Despite the encouraging achievements, there are still challenges that need to be addressed before AgGaS2 and its derivatives can gain equal popularity to other well-established nanocrystals such as chalcopyrite CuInS2, II-VI/III-V materials, and perovskites. Further research is needed to address issues related to synthesis scalability, long-term stability, and understanding the fundamental properties and behaviors of these materials.
First, historically, diamond-like compounds were studied extensively, but their NLO properties were largely overlooked until recent years. To fully exploit the potential of these materials, it is necessary to conduct a comprehensive structural identification of the available material systems. One effective approach is the incorporation of certain cations that exhibit the second-order Jahn–Teller effect, such as d0 transition-metal cations (e.g., Au+, Hf4+, Nb5+), and main-group cations that possess lone electron pairs (e.g., Pb2+, Sn2+, Bi3+, Sb3+, Te4+). These cations can introduce structural distortions and create favorable conditions for strong SHG responses. Furthermore, the modification of these compounds through appropriate alkali metal doping can be a promising strategy to increase the bandgap energy. This can lead to improved NLO properties, as well as enhanced stability and practicality for various applications. Advancements in high-throughput theoretical calculations have also been valuable in the search for high-quality NLO materials. These calculations can predict the NLO properties of materials based on their crystal structures and electronic properties, providing valuable guidance for experimental synthesis and characterization.
Second, the luminescence mechanism of AgGaS2-based nanocrystals requires further investigation and clarification. Some researchers have recorded large FWHMs of AgGaS2 nanocrystals (150–200 nm) [97] and attributed their luminescence to donor–acceptor pair recombination or free-to-bound recombination, which have been widely adopted for explaining the electronic transition of CuInS2. An ultra-large Stokes shift of 550–650 meV has also been observed in AgGaS2 nanocrystals [100], which is even more significant compared to the intensively studied CuInS2 nanocrystals. To understand the luminescence mechanism of AgGaS2 nanocrystals, several aspects should be addressed. On the one hand, if the donor–acceptor pair model is applicable, it is important to determine the locations of the donor and acceptor levels and investigate the dynamics of the transition using ultrafast spectroscopy in combination with theoretical modeling. This can provide insights into the energy levels involved in the luminescence processes. On the other hand, if the free-to-bound model is applicable, it is necessary to determine whether the recombination occurs from localized electronic defects to delocalized valence band (VB) states or from delocalized conduction band (CB) states to localized electronic defects. Understanding this process can shed light on the nature of the electronic defects and the fine structure of the electronic bands involved in the luminescence. To gain a deeper understanding of luminescence mechanisms, it would be beneficial to conduct systematic studies that investigate the excitation-dependent properties of emission. By carefully examining the emission spectra under different excitation conditions, valuable information about the intra-bandgap states and their involvement in the luminescence can be obtained.
Third, bulk AgGaS2 and AgGaS2 nanocrystals are expected to excel in novel applications. In the case of bulk AgGaS2, it is expected to excel in quantum imaging and sensing techniques in the mid-infrared range. Nondegenerate photon-pair generation through spontaneous parametric downconversion has been observed in AgGaS2 crystals. Researchers have reported the generation of idler photons in the mid-infrared spectral range above 6 µm in wavelength, accompanied by signal photons in the visible range [118]. This capability is significant considering the availability of single-photon detectors developed for infrared wavelengths up to 10 µm. The broad spectral bandwidth of photon-pair generation in AgGaS2 adds to its potential for various applications in quantum imaging and sensing [119]. On the other hand, AgGaS2 nanocrystals offer blue–green emission due to their suitable bandgap energy. This fluorescence in the range below 460 nm makes them suitable for a variety of lighting applications. Additionally, AgGaS2 nanocrystals can be used as shell materials in systems with smaller bandgaps, such as CuInS2 nanocrystals. By coating CuInS2 nanocrystals with AgGaS2, the resulting hybrid structures can exhibit multifunctional properties and enable a wider range of applications. Wet-chemical synthesis approaches for AgGaS2 nanocrystals hold promise for achieving multifunctionality and tailoring their properties for specific applications. These advancements open up exciting possibilities for the practical utilization of AgGaS2 in various fields.
Last but not the least, the environmental impact of the extraction of raw materials and the preparation of AgGaS2 and its derivatives deserves further investigation. As rare but critical metal elements, Ga and Ag have been widely used in electronic and optical devices in many forms besides AgGaS2, such as conductive silver pastes (for electrodes), copper–indium–gallium–chalcogenide materials (CIGS, for photovoltaic absorbers), silver sulfide (AgS, for near-infrared quantum dots), and so on. Certainly, the recovery and recycling of silver and gallium should be sufficiently prioritized to minimize their environmental impact.
Overall, the research on bulk and nanocrystal AgGaS2 driven by both scientific curiosity and application prospects is undoubtedly flourishing. Considering the rapid progress and growing interest in AgGaS2, it is reasonable to anticipate that this material will continue to play a significant and irreplaceable role in various fields, pushing the boundaries of NLO, optoelectronics, and beyond.

Author Contributions

Conceptualization, B.C.; writing—original draft preparation, G.X.; writing—review and editing, G.X. and B.C.; supervision, B.C.; funding acquisition, B.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (Grant No. 62205155) and the Natural Science Research Start-up Foundation of Recruiting Talents of Nanjing University of Posts and Telecommunications (Grant No. NY222104).

Data Availability Statement

The data are available upon request.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
NLONonlinear optics
OPOOptical parametric oscillation
SHGSecond harmonic generation
LIDTLaser-induced damage threshold
SOJTSecond-order Jahn–Teller
VECValence electron concentration
ICP-AESInductively coupled plasma atomic emission spectroscopy
FWHMFull width at half-maximum
LEDsLight-emitting diodes

References

  1. Sandroni, M.; Wegner, K.D.; Aldakov, D.; Reiss, P. Prospects of chalcopyrite-type nanocrystals for energy applications. ACS Energy Lett. 2017, 2, 1076–1088. [Google Scholar] [CrossRef]
  2. Yarema, O.; Yarema, M.; Wood, V. Tuning the composition of multicomponent semiconductor nanocrystals: The case of I–III–VI materials. Chem. Mater. 2018, 30, 1446–1461. [Google Scholar] [CrossRef]
  3. Berends, A.C.; Mangnus, M.J.; Xia, C.; Rabouw, F.T.; de Mello Donega, C. Optoelectronic properties of ternary I–III–VI2 semiconductor nanocrystals: Bright prospects with elusive origins. J. Phys. Chem. Lett. 2019, 10, 1600–1616. [Google Scholar] [CrossRef] [PubMed]
  4. Yang, L.; Zhang, S.; Xu, B.; Jiang, J.; Cai, B.; Lv, X.; Zou, Y.; Fan, Z.; Yang, H.; Zeng, H. I–III–VI Quantum Dots and Derivatives: Design, Synthesis, and Properties for Light-Emitting Diodes. Nano Lett. 2023, 23, 2443–2453. [Google Scholar] [CrossRef]
  5. Chen, B.; Zheng, W.; Chun, F.; Xu, X.; Zhao, Q.; Wang, F. Synthesis and hybridization of CuInS2 nanocrystals for emerging applications. Chem. Soc. Rev. 2023, 52, 8374–8409. [Google Scholar] [CrossRef] [PubMed]
  6. Kato, K.; Petrov, V.; Miyata, K. Refined Sellmeier equations for AgGaS2 down to 0.565  µm. Appl. Opt. 2023, 62, 6375–6379. [Google Scholar] [CrossRef] [PubMed]
  7. Luo, M.; Li, X.; Jiang, X.; Lin, Z.; Zhou, Z. The compatibly large nonlinear optical effect and high laser-induced damage threshold in a thiophosphate CsInP2S7 constructed with [P2S7]4− and [InS6]9−. Chin. Chem. Lett. 2024, 35, 109108. [Google Scholar] [CrossRef]
  8. Dong, L.; Zhang, S.; Gong, P.; Liang, F.; Lin, Z. AgIn5Se8: A defect diamond-like non-linear optical selenide. Inorg. Chem. Front. 2023, 10, 3248–3254. [Google Scholar] [CrossRef]
  9. Kinyaevskiy, I.O.; Koribut, A.V.; Grudtsyn, Y.V.; Ionin, M.V.; Klimachev, Y.M.; Yudin, N.N.; Ionin, A.A. Frequency conversion of femtosecond Ti: Sapphire laser pulse into 5–11 μm wavelength range with ZnGeP2, AgGaS2 and LiGaS2 crystals. Opt. Laser. Technol. 2025, 183, 112270. [Google Scholar] [CrossRef]
  10. Laksari, S.; Chahed, A.; Abbouni, N.; Benhelal, O.; Abbar, B. First-principles calculations of the structural, electronic and optical properties of CuGaS2 and AgGaS2. Comput. Mater. Sci 2006, 38, 223–230. [Google Scholar] [CrossRef]
  11. Abrahams, S.; Bernstein, J. Crystal structure of piezoelectric nonlinear-optic AgGaS2. J. Chem. Phys. 1973, 59, 1625–1629. [Google Scholar] [CrossRef]
  12. Rudysh, M.Y.; Ftomyn, N.; Shchepanskyi, P.; Myronchuk, G.; Popov, A.; Lemée, N.; Stadnyk, V.; Brik, M.; Piasecki, M. Electronic Structure, Optical, and Elastic Properties of AgGaS2 Crystal: Theoretical Study. Adv. Theory Simul. 2022, 5, 2200247. [Google Scholar] [CrossRef]
  13. Shannon, R.D. Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Crystallogr. Sect. A Cryst. Phys. Diffr. Theor. Gen. Crystallogr. 1976, 32, 751–767. [Google Scholar] [CrossRef]
  14. Huang, J.; Abudurusuli, A.; Yang, Z.; Pan, S. Emergent Mid-Infrared Nonlinear Optical Candidates with Targeted Balance Performances. Small 2024, 2409997. [Google Scholar] [CrossRef] [PubMed]
  15. Chen, B.; Zhu, S.; Zhao, B.; Lei, Y.; Wu, X.; Yuan, Z.; He, Z. Differential thermal analysis and crystal growth of AgGaS2. J. Cryst. Growth 2008, 310, 635–638. [Google Scholar] [CrossRef]
  16. Route, R.K.; Raymakers, R.; Feigelson, R. Preparation of large untwinned single crystals of AgGaS2. J. Cryst. Growth 1975, 29, 125–126. [Google Scholar] [CrossRef]
  17. Petit, J.; Bejet, M.; Daux, J.-C. Highly transparent AgGaS2 single crystals, a compound for mid-IR laser sources, using a combined static/dynamic vacuum annealing method. Mater. Chem. Phys. 2010, 119, 1–3. [Google Scholar] [CrossRef]
  18. Wu, J.; Huang, W.; Liu, H.-g.; He, Z.; Chen, B.; Zhu, S.; Zhao, B.; Lei, Y.; Zhou, X. Investigation of the thermal properties and crystal growth of the nonlinear optical crystals AgGaS2 and AgGaGeS4. Cryst. Growth Des. 2020, 20, 3140–3153. [Google Scholar] [CrossRef]
  19. Catella, G.C.; Burlage, D. Crystal growth and optical properties of AgGaS2 and AgGaSe2. MRS Bulletin 1998, 23, 28–36. [Google Scholar] [CrossRef]
  20. Exner, G.; Susner, M.A.; Murray, J.; Grigorov, A.; Siebenaller, R.; Conner, B.S.; Rowe, E.; Guha, S.; Petrov, V. Thermo-mechanical properties of BaGa2GeS(e)6 nonlinear optical crystals. Opt. Mater. Express 2023, 13, 1335–1344. [Google Scholar] [CrossRef]
  21. Sabanin, A.; Paterova, A. Prospects of using novel nonlinear crystals in SU(1,1) interferometry. Laser Phys. Lett. 2025, 22, 015205. [Google Scholar] [CrossRef]
  22. Suleimanova, D.Z.; Migal, E.A.; Badikov, D.V.; Potemkin, F.V. Advances in the development of few-cycle gigawatt-peak-power mid-IR optical parametric amplifiers pumped by Cr:Forsterite laser using non-oxide nonlinear crystals. Appl. Phys. Lett. 2024, 124, 161103. [Google Scholar] [CrossRef]
  23. Wu, K.; Pan, S. A review on structure-performance relationship toward the optimal design of infrared nonlinear optical materials with balanced performances. Coord. Chem. Rev. 2018, 377, 191–208. [Google Scholar] [CrossRef]
  24. Lin, H.; Wei, W.-B.; Chen, H.; Wu, X.-T.; Zhu, Q.-L. Rational design of infrared nonlinear optical chalcogenides by chemical substitution. Coord. Chem. Rev. 2020, 406, 213150. [Google Scholar] [CrossRef]
  25. Zaręba, J.K.; Nyk, M.; Samoć, M. Nonlinear Optical Properties of Emerging Nano-and Microcrystalline Materials. Adv. Opt. Mater. 2021, 9, 2100216. [Google Scholar] [CrossRef]
  26. Chen, H.; Ran, M.-Y.; Wei, W.-B.; Wu, X.-T.; Lin, H.; Zhu, Q.-L. A comprehensive review on metal chalcogenides with three-dimensional frameworks for infrared nonlinear optical applications. Coord. Chem. Rev. 2022, 470, 214706. [Google Scholar] [CrossRef]
  27. Hao, T.; Cen, Q.; Guan, S.; Li, W.; Dai, Y.; Zhu, N.; Li, M. Optoelectronic parametric oscillator. Light Sci. Appl. 2020, 9, 102. [Google Scholar] [CrossRef] [PubMed]
  28. Harmon, W.; Robben, K.; Cheatum, C.M. Adding a second AgGaS2 stage to Ti: Sapphire/BBO/AgGaS2 setups increases mid-infrared power twofold. Opt. Lett. 2023, 48, 4797–4800. [Google Scholar] [CrossRef] [PubMed]
  29. Ran, M.-Y.; Zhou, S.-H.; Wei, W.-B.; Li, B.-X.; Wu, X.-T.; Lin, H.; Zhu, Q.-L. Breaking Through the Trade-Off Between Wide Band Gap and Large SHG Coefficient in Mercury-Based Chalcogenides for IR Nonlinear Optical Application. Small 2024, 20, 2304563. [Google Scholar] [CrossRef] [PubMed]
  30. Chai, X.-D.; Li, M.-Z.; Lin, S.-J.; Chen, W.-F.; Jiang, X.-M.; Liu, B.-W.; Guo, G.-C. Cs4Zn5P6S18I2: The Largest Birefringence in Chalcohalide Achieved by Highly Polarizable Nonlinear Optical Functional Motifs. Small 2023, 19, 2303847. [Google Scholar] [CrossRef]
  31. Chemla, D.; Kupecek, P.; Robertson, D.; Smith, R. Silver thiogallate, a new material with potential for infrared devices. Opt. Commun. 1971, 3, 29–31. [Google Scholar] [CrossRef]
  32. Okorogu, A.; Mirov, S.; Lee, W.; Crouthamel, D.; Jenkins, N.; Dergachev, A.Y.; Vodopyanov, K.; Badikov, V. Tunable middle infrared downconversion in GaSe and AgGaS2. Opt. Commun. 1998, 155, 307–312. [Google Scholar] [CrossRef]
  33. Fan, Y.X.; Eckardt, R.; Byer, R.; Route, R.; Feigelson, R. AgGaS2 infrared parametric oscillator. Appl. Phys. Lett. 1984, 45, 313–315. [Google Scholar] [CrossRef]
  34. Vodopyanov, K.; Maffetone, J.; Zwieback, I.; Ruderman, W. AgGaS2 optical parametric oscillator continuously tunable from 3.9 to 11.3 μm. Appl. Phys. Lett. 1999, 75, 1204–1206. [Google Scholar] [CrossRef]
  35. Kildal, H.; Mikkelsen, J. The nonlinear optical coefficient, phasematching, and optical damage in the chalcopyrite AgGaSe2. Opt. Commun. 1973, 9, 315–318. [Google Scholar] [CrossRef]
  36. Boyd, G.; Buehler, E.; Storz, F. Linear and nonlinear optical properties of ZnGeP2 and CdSe. Appl. Phys. Lett. 1971, 18, 301–304. [Google Scholar] [CrossRef]
  37. Liang, F.; Kang, L.; Lin, Z.; Wu, Y.; Chen, C. Analysis and prediction of mid-IR nonlinear optical metal sulfides with diamond-like structures. Coord. Chem. Rev. 2017, 333, 57–70. [Google Scholar] [CrossRef]
  38. Hewage, N.W.; Viswanathan, G.; Yox, P.; Wu, K.; Kovnir, K.; Akopov, G. Synthesis of chiral high-entropy sulfides for non-linear optical applications. Inorg. Chem. Front. 2023, 10, 6613–6621. [Google Scholar] [CrossRef]
  39. Jomaa, M.; Mishra, V.; Mumbaraddi, D.; Sikdar, R.; Sarkar, D.; Sun, M.; Yao, J.; Michaelis, V.K.; Mar, A. Structure and Optical Properties of LixAg1–xGaSe2 and LixAg1–xInSe2. Inorg. Chem. 2023, 62, 7491–7502. [Google Scholar] [CrossRef]
  40. Ran, M.-Y.; Wang, A.Y.; Wei, W.-B.; Wu, X.-T.; Lin, H.; Zhu, Q.-L. Recent progress in the design of IR nonlinear optical materials by partial chemical substitution: Structural evolution and performance optimization. Coord. Chem. Rev. 2023, 481, 215059. [Google Scholar] [CrossRef]
  41. Kang, L.; Liang, F.; Jiang, X.; Lin, Z.; Chen, C. First-principles design and simulations promote the development of nonlinear optical crystals. Acc. Chem. Res. 2019, 53, 209–217. [Google Scholar] [CrossRef] [PubMed]
  42. Chen, H.; Ran, M.-Y.; Zhou, S.-H.; Wu, X.-T.; Lin, H. Ag2GeS3: A Diamond-Like Chalcogenide as an IR Nonlinear Optical Material with Outstanding Second-Harmonic Generation Response. Adv. Opt. Mater. 2024, 12, 2401100. [Google Scholar] [CrossRef]
  43. Han, Y.-X.; Hu, C.-L.; Chen, W.-T.; Mao, J.-G. Multi-step cation substitution facilitating the exploration of potential infrared nonlinear optical materials. Inorg. Chem. Front. 2024, 11, 5905–5912. [Google Scholar] [CrossRef]
  44. Lin, C.-S.; Zhou, A.-Y.; Cheng, W.-D.; Ye, N.; Chai, G.-L. Atom-resolved analysis of birefringence of nonlinear optical crystals by Bader charge integration. J. Phys. Chem. C 2019, 123, 31183–31189. [Google Scholar] [CrossRef]
  45. Yang, H.-D.; Ran, M.-Y.; Wei, W.-B.; Wu, X.-T.; Lin, H.; Zhu, Q.-L. Recent advances in IR nonlinear optical chalcogenides with well-balanced comprehensive performance. Mat. Today Phys. 2023, 35, 101127. [Google Scholar] [CrossRef]
  46. Bai, L.; Lin, Z.; Wang, Z.; Chen, C.; Lee, M.-H. Mechanism of linear and nonlinear optical effects of chalcopyrite AgGaX2 (X = S, Se, and Te) crystals. J. Chem. Phys. 2004, 120, 8772–8778. [Google Scholar] [CrossRef]
  47. Yang, S.; Lin, C.; Chen, K.; Fan, H.; Chen, J.; Fang, S.; Ye, N.; Luo, M. Trade-Off for Better Balanced Nonlinear Optical Performance with Disordered Si in ZnGeP2. Chem. Mater. 2022, 34, 11007–11013. [Google Scholar] [CrossRef]
  48. Yun, Y.; Xie, W.; Yang, Z.; Li, G.; Pan, S. Na+/Ag+ substitution induced birefringence enhancement from AgGaS2 to NaGaS2. J. Alloys Compd. 2022, 896, 163093. [Google Scholar] [CrossRef]
  49. Li, G.; Chu, Y.; Zhou, Z. From AgGaS2 to Li2ZnSiS4: Realizing impressive high laser damage threshold together with large second-harmonic generation response. Chem. Mater. 2018, 30, 602–606. [Google Scholar] [CrossRef]
  50. Chen, H.; Li, Y.-Y.; Li, B.; Liu, P.-F.; Lin, H.; Zhu, Q.-L.; Wu, X.-T. Salt-inclusion chalcogenide [Ba4Cl2][ZnGa4S10]: Rational design of an IR nonlinear optical material with superior comprehensive performance derived from AgGaS2. Chem. Mater. 2020, 32, 8012–8019. [Google Scholar] [CrossRef]
  51. Petrov, V.; Yelisseyev, A.; Isaenko, L.; Lobanov, S.; Titov, A.; Zondy, J.-J. Second harmonic generation and optical parametric amplification in the mid-IR with orthorhombic biaxial crystals LiGaS2 and LiGaSe2. Appl. Phys. B 2004, 78, 543–546. [Google Scholar] [CrossRef]
  52. Li, Y.-Y.; Liu, P.-F.; Wu, L.-M. Ba6Zn7Ga2S16: A wide band gap sulfide with phase-matchable infrared NLO properties. Chem. Mater. 2017, 29, 5259–5266. [Google Scholar] [CrossRef]
  53. Zhou, H.M.; Xiong, L.; Chen, L.; Wu, L.M. Dislocations that decrease size mismatch within the lattice leading to ultrawide band gap, large second-order susceptibility, and high nonlinear optical performance of AgGaS2. Angew. Chem. Int. Ed. 2019, 58, 9979–9983. [Google Scholar] [CrossRef]
  54. Li, G.; Wu, K.; Liu, Q.; Yang, Z.; Pan, S. Na2ZnGe2S6: A new infrared nonlinear optical material with good balance between large second-harmonic generation response and high laser damage threshold. J. Am. Chem. Soc. 2016, 138, 7422–7428. [Google Scholar] [CrossRef] [PubMed]
  55. Li, S.-F.; Jiang, X.-M.; Liu, B.-W.; Yan, D.; Zeng, H.-Y.; Guo, G.-C. Strong infrared nonlinear optical efficiency and high laser damage threshold realized in quaternary alkali metal sulfides Na2Ga2MS6 (M = Ge, Sn) containing mixed nonlinear optically active motifs. Inorg. Chem. 2018, 57, 6783–6786. [Google Scholar] [CrossRef] [PubMed]
  56. Xing, W.; Wang, N.; Tang, C.; Li, C.; Lin, Z.; Yao, J.; Yin, W.; Kang, B. From AgGaS2 to AgHgPS4: Vacancy defects and highly distorted HgS4 tetrahedra double-induced remarkable second-harmonic generation response. J. Mater. Chem. C 2021, 9, 1062–1068. [Google Scholar] [CrossRef]
  57. Gao, L.; Xu, J.; Tian, X.; Zhang, B.; Wu, X.; Wu, K. AgGaSe2-inspired nonlinear optical materials: Tetrel selenides of alkali metals and mercury. Chem. Mater. 2022, 34, 5991–5998. [Google Scholar] [CrossRef]
  58. Zhang, X.; Wu, H.; Hu, Z.; Wang, J.; Wu, Y.; Yu, H. AllHgMIVS4 (All = Sr, Ba, MIV = Si, Ge): A Series of Materials with Large Second Harmonic Generation Response and Wide Band Gaps. Adv. Opt. Mater. 2024, 12, 2301735. [Google Scholar] [CrossRef]
  59. Grechenkov, J.; Gopejenko, A.; Bocharov, D.; Isakoviča, I.; Popov, A.I.; Brik, M.G.; Piskunov, S. Ab Initio Modeling of CuGa1−xInxS2, CuGaS2(1−x)Se2x and Ag1−xCuxGaS2 Chalcopyrite Solid Solutions for Photovoltaic Applications. Energies 2023, 16, 4823. [Google Scholar] [CrossRef]
  60. Li, J.; Yao, W.-D.; Li, J.-N.; Li, X.-H.; Liu, W.; Guo, S.-P. Partial substitution induced structural transformation and enhanced nonlinear optical properties of Na2GaxIn6-xSe10 (x = 3, 3.76). Mater. Today Phys. 2023, 32, 101007. [Google Scholar] [CrossRef]
  61. Tang, G.; He, Y.; Liao, H.; Zhang, S.; Chu, X.; Si, R.; Jin, Z.; Du, Y.; Wang, C.; Chen, B.; et al. Crystal Growth, Characterization, and Properties of Nonlinear Optical Crystals of Li0.5Ag0.5GaTe2 for Mid-Infrared Applications. ACS Appl. Mater. Inter. 2024, 16, 62411–62420. [Google Scholar] [CrossRef]
  62. Tang, J.; Yao, A.; Wu, Q.; Tang, C.; Wu, J.; Kang, L.; Xia, M.; Yin, W.; Kang, B.; Deng, J. AAg2MS4 (A = Na, K, Rb; M = P, As) Nonlinear Optical Series with Unique Cation-Regulated Highly Anisotropic [AgS4] Hypertetrahedral Layered Framework. Adv. Opt. Mater. 2024, 12, 2401126. [Google Scholar] [CrossRef]
  63. Adhikary, A.; Yaghoobnejad Asl, H.; Sandineni, P.; Balijapelly, S.; Mohapatra, S.; Khatua, S.; Konar, S.; Gerasimchuk, N.; Chernatynskiy, A.V.; Choudhury, A. Unusual atmospheric water trapping and water induced reversible restacking of 2D gallium sulfide layers in NaGaS2 formed by supertetrahedral building unit. Chem. Mater. 2020, 32, 5589–5603. [Google Scholar] [CrossRef]
  64. Klepov, V.V.; Berseneva, A.A.; Pace, K.A.; Kocevski, V.; Sun, M.; Qiu, P.; Wang, H.; Chen, F.; Besmann, T.M.; zur Loye, H.C. NaGaS2: An Elusive Layered Compound with Dynamic Water Absorption and Wide-Ranging Ion-Exchange Properties. Angew. Chem. Int. Ed. 2020, 59, 10836–10841. [Google Scholar] [CrossRef]
  65. Zou, G.; Ye, N.; Huang, L.; Lin, X. Alkaline-alkaline earth fluoride carbonate crystals ABCO3F (A = K, Rb, Cs; B = Ca, Sr, Ba) as nonlinear optical materials. J. Am. Chem. Soc. 2011, 133, 20001–20007. [Google Scholar] [CrossRef] [PubMed]
  66. Liu, B.-W.; Zeng, H.-Y.; Zhang, M.-J.; Fan, Y.-H.; Guo, G.-C.; Huang, J.-S.; Dong, Z.-C. Syntheses, structures, and nonlinear-optical properties of metal sulfides Ba2Ga8MS16 (M = Si, Ge). Inorg. Chem. 2015, 54, 976–981. [Google Scholar] [CrossRef] [PubMed]
  67. Li, Z.; Zhang, S.; Huang, Z.; Zhao, L.-D.; Uykur, E.; Xing, W.; Lin, Z.; Yao, J.; Wu, Y. Molecular construction from AgGaS2 to CuZnPS4: Defect-induced second harmonic generation enhancement and cosubstitution-driven band gap enlargement. Chem. Mater. 2020, 32, 3288–3296. [Google Scholar] [CrossRef]
  68. Nguyen, V.; Ji, B.; Wu, K.; Zhang, B.; Wang, J. Unprecedented mid-infrared nonlinear optical materials achieved by crystal structure engineering, a case study of (KX)P2S6 (X = Sb, Bi, Ba). Chem. Sci. 2022, 13, 2640–2648. [Google Scholar] [CrossRef] [PubMed]
  69. Feng, J.H.; Hu, C.L.; Xu, X.; Li, B.X.; Zhang, M.J.; Mao, J.G. AgGa2PS6: A New Mid-Infrared Nonlinear Optical Material with a High Laser Damage Threshold and a Large Second Harmonic Generation Response. Chem. Euro. J. 2017, 23, 10978–10982. [Google Scholar] [CrossRef] [PubMed]
  70. Huang, X.; Yang, S.H.; Li, X.H.; Liu, W.; Guo, S.P. Eu2P2S6: The First Rare-Earth Chalcogenophosphate Exhibiting Large Second-Harmonic Generation Response and High Laser-Induced Damage Threshold. Angew. Chem. Int. Ed. 2022, 61, e202206791. [Google Scholar] [CrossRef] [PubMed]
  71. Guo, S.-P.; Chi, Y.; Guo, G.-C. Recent achievements on middle and far-infrared second-order nonlinear optical materials. Coord. Chem. Rev. 2017, 335, 44–57. [Google Scholar] [CrossRef]
  72. Li, C.; Meng, X.; Li, Z.; Yao, J. Hg-based chalcogenides: An intriguing class of infrared nonlinear optical materials. Coord. Chem. Rev. 2022, 453, 214328. [Google Scholar] [CrossRef]
  73. Guo, Y.; Liang, F.; Yao, J.; Lin, Z.; Yin, W.; Wu, Y.; Chen, C. Nonbonding electrons driven strong SHG effect in Hg2GeSe4: Experimental and theoretical investigations. Inorg. Chem. 2018, 57, 6795–6798. [Google Scholar] [CrossRef] [PubMed]
  74. Li, J.; Li, X.-H.; Yao, W.-D.; Liu, W.; Guo, S.-P. Three-in-One Strategy Constructing the First High-Performance Nonlinear Optical Sulfide Crystallizing with the P43 Space Group. Small 2023, 19, 2303090. [Google Scholar] [CrossRef]
  75. Lou, X.-Y.; Jiang, X.-M.; Liu, B.-W.; Guo, G.-C. Excellent Nonlinear Optical M[M4Cl][Ga11S20] (M = A/Ba, A = K, Rb) Achieved by Unusual Cationic Substitution Strategy. Small 2024, 20, 2305711. [Google Scholar] [CrossRef] [PubMed]
  76. Zhong, J.; Zhang, B.; Pan, F.; Lin, Z. Thiotetrelates Li2ZnXS4 (X = Si, Ge, and Sn) As Potential Li-Ion Solid-State Electrolytes. ACS Appl. Mater. Inter. 2022, 14, 9203–9211. [Google Scholar] [CrossRef] [PubMed]
  77. Berseneva, A.A.; Masachchi, L.W.; Jacobsohn, L.G.; zur Loye, H.-C. Tunable Salt-Inclusion Chalcogenides for Ion Exchange, Photoluminescence, and Scintillation. Chem. Mater. 2023, 35, 1417–1431. [Google Scholar] [CrossRef]
  78. Yue, Q.-G.; Wei, W.-B.; Chen, H.; Wu, X.-T.; Lin, H.; Zhu, Q.-L. Salt-inclusion chalcogenides: An emerging class of IR nonlinear optical materials. Dalton Trans. 2020, 49, 14338–14343. [Google Scholar] [CrossRef] [PubMed]
  79. Liu, B.-W.; Jiang, X.-M.; Zeng, H.-Y.; Guo, G.-C. [ABa2Cl][Ga4S8](A = Rb, Cs): Wide-spectrum nonlinear optical materials obtained by polycation-substitution-induced nonlinear optical (NLO)-functional motif ordering. J. Am. Chem. Soc. 2020, 142, 10641–10645. [Google Scholar] [CrossRef] [PubMed]
  80. Chen, W.-F.; Jiang, X.-M.; Liu, B.-W.; Guo, G.-C. Wide band gaps in nonlinear optical materials AMgAgGa6S11 (A = K, Rb) achieved by the incorporation of dual s-block elements in Ag-based sulfide. Mater. Today Phys. 2023, 37, 101199. [Google Scholar] [CrossRef]
  81. Zhou, W.; Wu, J.; Liu, W.; Guo, S.-P. Ag-based chalcogenides and derivatives as promising infrared nonlinear optical materials. Coord. Chem. Rev. 2023, 477, 214950. [Google Scholar] [CrossRef]
  82. Wu, H.; Li, X.; Cheng, Y.; Xiao, Y.; Wu, Q.; Lin, H.; Xu, J.; Wang, Y. The synergistic role of double vacancies within AgGaS2 nanocrystals in carrier separation and transfer for efficient photocatalytic hydrogen evolution. Catal. Sci. Technol. 2019, 9, 5838–5844. [Google Scholar] [CrossRef]
  83. Xie, X.; Zhao, J.; Lin, O.; Yin, Z.; Li, X.; Zhang, Y.; Tang, A. Narrow-Bandwidth Blue-Emitting Ag–Ga–Zn–S Semiconductor Nanocrystals for Quantum-Dot Light-Emitting Diodes. J. Phys. Chem. Lett. 2022, 13, 11857–11863. [Google Scholar] [CrossRef] [PubMed]
  84. Li, X.; Tong, X.; Xia, L.; Zhao, H.; Luo, J.; Li, Z.; Wang, Z.M. Modulating Eco-friendly Colloidal AgGaS2 Quantum Dots for Highly Efficient Photodetection and Image Sensing via Direct Growth of Ternary AgInS2 Shell. Small 2024, 20, 2404261. [Google Scholar] [CrossRef] [PubMed]
  85. Wang, X.; Yang, J.; Liu, X.; Zhong, H.; Zhang, Z.; Luo, Q.; Chen, D.; Liu, J.; Lin, H. Indium-free silver thiogallate nanoflake-clusters grown on carbon nitride edges for hydrogen peroxide photosynthesis with enhanced activity and stability. J. Colloid Interface Sci. 2025, 680, 581–594. [Google Scholar] [CrossRef]
  86. Lin, O.; Wang, L.; Xie, X.; Wang, S.; Feng, Y.; Xiao, J.; Zhang, Y.; Tang, A. Seed-mediated growth synthesis and tunable narrow-band luminescence of quaternary Ag–In–Ga–S alloyed nanocrystals. Nanoscale 2024, 16, 4591–4599. [Google Scholar] [CrossRef] [PubMed]
  87. Alivisatos, A.P. Semiconductor clusters, nanocrystals, and quantum dots. Science 1996, 271, 933–937. [Google Scholar] [CrossRef]
  88. Tozawa, M.; Ofuji, S.; Tanaka, M.; Akiyoshi, K.; Kameyama, T.; Yamamoto, T.; Motomura, G.; Fujisaki, Y.; Uematsu, T.; Kuwabata, S.; et al. Spectrally Narrow Blue-Light Emission from Nonstoichiometric AgGaS2 Quantum Dots for Application to Light-Emitting Diodes. ACS Appl. Mater. Inter. 2024, 16, 68169–68180. [Google Scholar] [CrossRef]
  89. Leach, A.D.; Macdonald, J.E. Optoelectronic properties of CuInS2 nanocrystals and their origin. J. Phys. Chem. Lett. 2016, 7, 572–583. [Google Scholar] [CrossRef]
  90. Adhikari, S.D.; Dutta, A.; Prusty, G.; Sahu, P.; Pradhan, N. Symmetry break and seeded 2D anisotropic growth in ternary CuGaS2 nanocrystals. Chem. Mater. 2017, 29, 5384–5393. [Google Scholar] [CrossRef]
  91. Zhong, H.; Wang, Z.; Bovero, E.; Lu, Z.; Van Veggel, F.C.; Scholes, G.D. Colloidal CuInSe2 nanocrystals in the quantum confinement regime: Synthesis, optical properties, and electroluminescence. J. Phys. Chem. C 2011, 115, 12396–12402. [Google Scholar] [CrossRef]
  92. Hu, J.; Lu, Q.; Tang, K.; Qian, Y.; Zhou, G.; Liu, X. Solvothermal reaction route to nanocrystalline semiconductors AgMS2 (M = Ga, In). Chem. Commun. 1999, 12, 1093–1094. [Google Scholar] [CrossRef]
  93. Junqing, H.; Bin, D.; Kaibin, T.; Qingyi, L.; Rongrong, J.; Yitai, Q. Hydrothermal preparation and characterization of nanocrystalline: Silver gallium sulfides. Solid State Sci. 2001, 3, 275–278. [Google Scholar] [CrossRef]
  94. Fan, C.-M.; Regulacio, M.D.; Ye, C.; Lim, S.H.; Zheng, Y.; Xu, Q.-H.; Xu, A.-W.; Han, M.-Y. Colloidal synthesis and photocatalytic properties of orthorhombic AgGaS2 nanocrystals. Chem. Commun. 2014, 50, 7128–7131. [Google Scholar] [CrossRef]
  95. Paderick, S.; Kessler, M.; Hurlburt, T.J.; Hughes, S.M. Synthesis and characterization of AgGaS2 nanoparticles: A study of growth and fluorescence. Chem. Commun. 2018, 54, 62–65. [Google Scholar] [CrossRef] [PubMed]
  96. Azhniuk, Y.; Lopushanska, B.; Selyshchev, O.; Havryliuk, Y.; Pogodin, A.; Kokhan, O.; Ehm, A.; Lopushansky, V.; Studenyak, I.; Zahn, D.R. Synthesis and Optical Properties of Ag–Ga–S Quantum Dots. Phys. Status Solidi B 2022, 259, 2100349. [Google Scholar] [CrossRef]
  97. Uematsu, T.; Doi, T.; Torimoto, T.; Kuwabata, S. Preparation of luminescent AgInS2–AgGaS2 solid solution nanoparticles and their optical properties. J. Phys. Chem. Lett. 2010, 1, 3283–3287. [Google Scholar] [CrossRef]
  98. Song, J.; Zhang, Y.; Dai, Y.; Hu, J.; Zhu, L.; Xu, X.; Yu, Y.; Li, H.; Yao, B.; Zhou, H. Polyelectrolyte-Mediated Nontoxic AgGaxIn1–xS2 QDs/Low-Density Lipoprotein Nanoprobe for Selective 3D Fluorescence Imaging of Cancer Stem Cells. ACS Appl. Mater. Inter. 2019, 11, 9884–9892. [Google Scholar] [CrossRef]
  99. Li, T.; Shang, D.; Liu, C.; Li, R.; Bai, F.; Bai, T.; Xing, S. Colloidal synthesis of AgGa (S1−xSex)2 solid solution nanocrystals with composition-dependent crystal phase for efficient photocatalytic degradation of methyl violet. CrystEngComm 2022, 24, 4540–4545. [Google Scholar] [CrossRef]
  100. Kim, J.-H.; Kim, B.-Y.; Jang, E.-P.; Yoon, S.-Y.; Kim, K.-H.; Do, Y.R.; Yang, H. Synthesis of widely emission-tunable Ag–Ga–S and its quaternary derivative quantum dots. Chem. Eng. J. 2018, 347, 791–797. [Google Scholar] [CrossRef]
  101. Bai, T.; Wang, X.; Dong, Y.; Xing, S.; Shi, Z.; Feng, S. One-pot synthesis of high-quality AgGaS2/ZnS-based photoluminescent nanocrystals with widely tunable band gap. Inorg. Chem. Front. 2020, 59, 5975–5982. [Google Scholar] [CrossRef]
  102. Huang, G.; Huang, Y.; Liu, Z.; Wei, J.; Zhu, Q.; Jiang, G.; Jin, X.; Li, Q.; Li, F. White light-emitting diodes based on quaternary Ag–In-Ga-S quantum dots and their influences on melatonin suppression index. J. Lumin. 2021, 233, 117903. [Google Scholar] [CrossRef]
  103. Kameyama, T.; Kishi, M.; Miyamae, C.; Sharma, D.K.; Hirata, S.; Yamamoto, T.; Uematsu, T.; Vacha, M.; Kuwabata, S.; Torimoto, T. Wavelength-tunable band-edge photoluminescence of nonstoichiometric Ag–In–S nanoparticles via Ga3+ doping. ACS Appl. Mater. Inter. 2018, 10, 42844–42855. [Google Scholar] [CrossRef]
  104. Hoisang, W.; Uematsu, T.; Torimoto, T.; Kuwabata, S. Luminescent Quaternary Ag(InxGa1–x)S2/GaSy Core/Shell Quantum Dots Prepared Using Dithiocarbamate Compounds and Photoluminescence Recovery via Post Treatment. Inorg. Chem. 2021, 60, 13101–13109. [Google Scholar] [CrossRef] [PubMed]
  105. Li, X.; Tong, X.; Yue, S.; Liu, C.; Channa, A.I.; You, Y.; Wang, R.; Long, Z.; Zhang, Z.; Zhao, Z. Rational design of colloidal AgGaS2/CdSeS core/shell quantum dots for solar energy conversion and light detection. Nano Energy 2021, 89, 106392. [Google Scholar] [CrossRef]
  106. LaMer, V.K.; Dinegar, R.H. Theory, production and mechanism of formation of monodispersed hydrosols. J. Am. Chem. Soc. 1950, 72, 4847–4854. [Google Scholar] [CrossRef]
  107. Alves Machado Filho, M.; Hsiao, C.-L.; dos Santos, R.B.; Hultman, L.; Birch, J.; Gueorguiev, G.K. Self-Induced Core–Shell InAlN Nanorods: Formation and Stability Unraveled by Ab Initio Simulations. ACS Nanosci. Au 2023, 3, 84–93. [Google Scholar] [CrossRef]
  108. Filho, M.A.M.; Farmer, W.; Hsiao, C.-L.; dos Santos, R.B.; Hultman, L.; Birch, J.; Ankit, K.; Gueorguiev, G.K. Density Functional Theory-Fed Phase Field Model for Semiconductor Nanostructures: The Case of Self-Induced Core–Shell InAlN Nanorods. Cryst. Growth. Des. 2024, 24, 4717–4727. [Google Scholar] [CrossRef] [PubMed]
  109. Chen, B.; Li, D.; Wang, F. InP quantum dots: Synthesis and lighting applications. Small 2020, 16, 2002454. [Google Scholar] [CrossRef]
  110. Xie, X.; Zhao, J.; Lin, O.; Niu, W.; Li, Y.; Wang, L.; Li, L.; Yin, Z.; Li, X.; Zhang, Y.; et al. One-Pot Synthesis of Color-Tunable Narrow-Bandwidth Ag–In–Ga–Zn–S Semiconductor Nanocrystals for Quantum-Dot Light-Emitting Diodes. Nano Lett. 2024, 24, 9683–9690. [Google Scholar] [CrossRef] [PubMed]
  111. Ji, B.; Koley, S.; Slobodkin, I.; Remennik, S.; Banin, U. ZnSe/ZnS core/shell quantum dots with superior optical properties through thermodynamic shell growth. Nano Lett. 2020, 20, 2387–2395. [Google Scholar] [CrossRef] [PubMed]
  112. Won, Y.-H.; Cho, O.; Kim, T.; Chung, D.-Y.; Kim, T.; Chung, H.; Jang, H.; Lee, J.; Kim, D.; Jang, E. Highly efficient and stable InP/ZnSe/ZnS quantum dot light-emitting diodes. Nature 2019, 575, 634–638. [Google Scholar] [CrossRef] [PubMed]
  113. Uematsu, T.; Tepakidareekul, M.; Hirano, T.; Torimoto, T.; Kuwabata, S. Facile High-Yield Synthesis of Ag–In–Ga–S Quaternary Quantum Dots and Coating with Gallium Sulfide Shells for Narrow Band-Edge Emission. Chem. Mater. 2023, 35, 1094–1106. [Google Scholar] [CrossRef]
  114. Zhao, J.; Chen, B.; Wang, F. Shedding light on the role of misfit strain in controlling core–shell nanocrystals. Adv. Mater. 2020, 32, 2004142. [Google Scholar] [CrossRef]
  115. Berends, A.C.; Van Der Stam, W.; Hofmann, J.P.; Bladt, E.; Meeldijk, J.D.; Bals, S.; de Mello Donega, C. Interplay between surface chemistry, precursor reactivity, and temperature determines outcome of ZnS shelling reactions on CuInS2 nanocrystals. Chem. Mater. 2018, 30, 2400–2413. [Google Scholar] [CrossRef]
  116. Chuang, P.-H.; Lin, C.C.; Liu, R.-S. Emission-tunable CuInS2/ZnS quantum dots: Structure, optical properties, and application in white light-emitting diodes with high color rendering index. ACS Appl. Mater. Inter. 2014, 6, 15379–15387. [Google Scholar] [CrossRef] [PubMed]
  117. Jara, D.H.; Stamplecoskie, K.G.; Kamat, P.V. Two distinct transitions in CuxInS2 quantum dots. Bandgap versus sub-bandgap excitations in copper-deficient structures. J. Phys. Chem. Lett. 2016, 7, 1452–1459. [Google Scholar] [CrossRef] [PubMed]
  118. Kumar, M.; Kumar, P.; Vega, A.; Weissflog, M.A.; Pertsch, T.; Setzpfandt, F. Mid-infrared photon pair generation in AgGaS2. Appl. Phys. Lett. 2021, 119, 244001. [Google Scholar] [CrossRef]
  119. Trovatello, C.; Marini, A.; Xu, X.; Lee, C.; Liu, F.; Curreli, N.; Manzoni, C.; Dal Conte, S.; Yao, K.; Ciattoni, A. Optical parametric amplification by monolayer transition metal dichalcogenides. Nat. Photonics 2021, 15, 6–10. [Google Scholar] [CrossRef]
Figure 2. (a) Pseudo-binary phase diagram of the Ag2S–Ga2S3 system. Reproduced with permission [15]. Copyright 2007, Elsevier B.V. (b) Photograph of a AgGaS2 single crystal grown along the [001] direction with full size of 40 mm × 90 mm. (c) Temperature-dependent thermal expansion coefficients of a AgGaS2 crystal in the direction of the a and c axes, respectively. Reproduced with permission [18]. Copyright 2020, American Chemical Society. (d) Photograph of a AgGaS2 crystal annealed by alternant static and dynamic vacuum annealing. Note that the sample is 4 mm in thickness. Reprinted with permission [17]. Copyright 2010 Elsevier B.V.
Figure 2. (a) Pseudo-binary phase diagram of the Ag2S–Ga2S3 system. Reproduced with permission [15]. Copyright 2007, Elsevier B.V. (b) Photograph of a AgGaS2 single crystal grown along the [001] direction with full size of 40 mm × 90 mm. (c) Temperature-dependent thermal expansion coefficients of a AgGaS2 crystal in the direction of the a and c axes, respectively. Reproduced with permission [18]. Copyright 2020, American Chemical Society. (d) Photograph of a AgGaS2 crystal annealed by alternant static and dynamic vacuum annealing. Note that the sample is 4 mm in thickness. Reprinted with permission [17]. Copyright 2010 Elsevier B.V.
Nanomaterials 15 00147 g002
Figure 3. Structural evolution of diamond-like chalcogenides from II–VI (ZnS) to I–III–VI2 (AgGaS2), I2–II–IV–VI4 (Cu2CdSnS4). Reproduced with permission [41]. Copyright 2019, American Chemical Society.
Figure 3. Structural evolution of diamond-like chalcogenides from II–VI (ZnS) to I–III–VI2 (AgGaS2), I2–II–IV–VI4 (Cu2CdSnS4). Reproduced with permission [41]. Copyright 2019, American Chemical Society.
Nanomaterials 15 00147 g003
Figure 4. Crystal structures of (a) (Li/Ag)GaS2, (b) NaGaS2, (c) Ba2Ga8Ge16, and (d) CuZnPS4 showing mixed-anion polyhedrons arranged in diamond-like configurations.
Figure 4. Crystal structures of (a) (Li/Ag)GaS2, (b) NaGaS2, (c) Ba2Ga8Ge16, and (d) CuZnPS4 showing mixed-anion polyhedrons arranged in diamond-like configurations.
Nanomaterials 15 00147 g004
Figure 5. (a) Crystal structure of AgHgPS4. (b) Structural evolution from AgGaS2 to AgHgPS4. Reproduced with permission [56]. Copyright 2021, The Royal Society of Chemistry. (c) Crystal structure of Li2HgSnSe4. (d) Structural evolution from AgGaSe2 to Li2HgSnSe4. Reproduced with permission [57]. Copyright 2022, American Chemical Society.
Figure 5. (a) Crystal structure of AgHgPS4. (b) Structural evolution from AgGaS2 to AgHgPS4. Reproduced with permission [56]. Copyright 2021, The Royal Society of Chemistry. (c) Crystal structure of Li2HgSnSe4. (d) Structural evolution from AgGaSe2 to Li2HgSnSe4. Reproduced with permission [57]. Copyright 2022, American Chemical Society.
Nanomaterials 15 00147 g005
Figure 6. (a) Crystal structure of Na2Ga2MS6 (M = Ge or Sn). (b) Crystal structure of Li2ZnSiS4. Reproduced with permission [76]. Copyright 2022, American Chemical Society. (c) Crystal structure of [Ba4Cl2][ZnGa4S10]. Reproduced with permission [50]. Copyright 2020, American Chemical Society.
Figure 6. (a) Crystal structure of Na2Ga2MS6 (M = Ge or Sn). (b) Crystal structure of Li2ZnSiS4. Reproduced with permission [76]. Copyright 2022, American Chemical Society. (c) Crystal structure of [Ba4Cl2][ZnGa4S10]. Reproduced with permission [50]. Copyright 2020, American Chemical Society.
Nanomaterials 15 00147 g006
Figure 7. (a) XRD pattern and (b) TEM images of AgGaS2 nanocrystals prepared using dodecanethiol (DDT), hexadecanethiol (HDT), oleylamine (OM), and hexadecylamine (HAD). Reproduced with permission [94]. Copyright 2014, The Royal Society of Chemistry.
Figure 7. (a) XRD pattern and (b) TEM images of AgGaS2 nanocrystals prepared using dodecanethiol (DDT), hexadecanethiol (HDT), oleylamine (OM), and hexadecylamine (HAD). Reproduced with permission [94]. Copyright 2014, The Royal Society of Chemistry.
Nanomaterials 15 00147 g007
Figure 8. (a) Scheme of the synthesis of AgGaS2 nanocrystals through a cation-exchange approach. (b) TEM images and high-resolution TEM images of AgGaxIn1-XS2 nanocrystals (top panel) and AgGaS2 nanocrystals (bottom), respectively. Scale bars are 10 nm and 3 nm for TEM images and high-resolution TEM images. (c) Absorption and emission spectra of nanocrystals with [Ca]/[In] ratios of (i) 0.85, (ii) 1.56, (iii) 2.88, and (iv) 3.73 in precursors, respectively. Reproduced with permission [98]. Copyright 2019, American Chemical Society.
Figure 8. (a) Scheme of the synthesis of AgGaS2 nanocrystals through a cation-exchange approach. (b) TEM images and high-resolution TEM images of AgGaxIn1-XS2 nanocrystals (top panel) and AgGaS2 nanocrystals (bottom), respectively. Scale bars are 10 nm and 3 nm for TEM images and high-resolution TEM images. (c) Absorption and emission spectra of nanocrystals with [Ca]/[In] ratios of (i) 0.85, (ii) 1.56, (iii) 2.88, and (iv) 3.73 in precursors, respectively. Reproduced with permission [98]. Copyright 2019, American Chemical Society.
Nanomaterials 15 00147 g008
Figure 9. (a) Normalized photoluminescence spectra (λex = 470 nm) of AgGa1−yInyS2 (y = 0.2~1.0) nanocrystals. Reproduced with permission [97]. Copyright 2010, American Chemical Society. (b) Normalized photoluminescence spectra of AgGaS2 and AgGaS2: Zn nanocrystals. Insets: corresponding photos of AgGaS2 and AgGaS2: Zn nanocrystals under natural light and 365 nm ultraviolet light. Reproduced with permission [83]. Copyright 2022, American Chemical Society.
Figure 9. (a) Normalized photoluminescence spectra (λex = 470 nm) of AgGa1−yInyS2 (y = 0.2~1.0) nanocrystals. Reproduced with permission [97]. Copyright 2010, American Chemical Society. (b) Normalized photoluminescence spectra of AgGaS2 and AgGaS2: Zn nanocrystals. Insets: corresponding photos of AgGaS2 and AgGaS2: Zn nanocrystals under natural light and 365 nm ultraviolet light. Reproduced with permission [83]. Copyright 2022, American Chemical Society.
Nanomaterials 15 00147 g009
Figure 10. (a) Photographs and photoluminescence spectra of AgGaS2@ZnS nanocrystals with varied Ag/Ga or In/Ga ratios. Reproduced with permission [101]. Copyright 2020, American Chemical Society. (b) Photoluminescence spectra of AgGa0.6In0.4S2 and AgGa0.6In0.4S2@GaSx nanocrystals. Reproduced with permission [103]. Copyright 2018, American Chemical Society. (c) Absorption and photoluminescence spectra of AgGaS2 and AgGaS2@CdSeS nanocrystals, respectively. Reproduced with permission [105]. Copyright 2021, Elsevier Ltd.
Figure 10. (a) Photographs and photoluminescence spectra of AgGaS2@ZnS nanocrystals with varied Ag/Ga or In/Ga ratios. Reproduced with permission [101]. Copyright 2020, American Chemical Society. (b) Photoluminescence spectra of AgGa0.6In0.4S2 and AgGa0.6In0.4S2@GaSx nanocrystals. Reproduced with permission [103]. Copyright 2018, American Chemical Society. (c) Absorption and photoluminescence spectra of AgGaS2 and AgGaS2@CdSeS nanocrystals, respectively. Reproduced with permission [105]. Copyright 2021, Elsevier Ltd.
Nanomaterials 15 00147 g010
Table 1. NLO parameters of AgGaS2 and typical AgGaS2-derived crystals with diamond-like structures.
Table 1. NLO parameters of AgGaS2 and typical AgGaS2-derived crystals with diamond-like structures.
TypeCompoundSpace GroupEg (eV)Band TypesIR Cutoff (μm)Δ[email protected] μmSHG dij (pm/V)LIDT
(×AgGaS2)
Ref
Commercially availableAgGaS2I-42d2.7Direct0.45–130.053d36 = 13.91.0[34]
AgGaSe2I-42d1.8Direct0.76–17N.A.d36 = 33.0N.A.[46]
ZnGeP2I-42d1.74Direct0.74–120.046deff = 75.0N.A.[47]
Bandgap modificationNaGaS2I-42d3.9Direct0.31–13.30.094d36 = 13.21.1[48]
Li2ZnSiS4Pna213.9Direct25N.A.d33 = 18.910.0[49]
[Ba4Cl2]
[ZnGa4S10]
I-43.8Direct0.29–13.7[email protected] μmd14 = 14.951.0[50]
LiGaS2Pna213.7Direct11.60.014d33 = 11.211.0[51]
Ba6Zn7Ga2S16R33.5DirectN.A.0.036d11 = 6.128.0[52]
AgGaS2:LiI-42d3.4DirectN.A.N.A.d36 = 20.68.6[53]
Na2ZnGe2S6Cc3.2Direct0.38–220.026d33 = −5.36.0[54]
Na2Ga2GeS6Fdd23.1DirectN.A.N.A.deff = 11.218.1[55]
Ba2Ga8GeS16P63mc3.0Direct0.42–20N.A.deff = 26.022.0
CuZnPS4I-42m3.0Indirect16.50.07d14 = 15.96.0
SHG enhancementNa2Ga2SnS6Fdd22.7DirectN.A.N.A.deff = 13.117.9[55]
AgHgPS4Pn2.6Indirect12[email protected] μmd11 = −31.1N.A.[56]
Li2HgGeSe4Pna211.7DirectN.A.0.042–0.074d33 = −90.23.0[57]
Li2HgSnSe4Pna211.6DirectN.A.d33 = −1043.5
Na2Hg3Ge2Se8P-4c21.3DirectN.A.0.096–0.172d36 = 87.64.5
Na2Hg3Sn2Se8P-4c21.2DirectN.A.d36 = 96.43.0
Table 2. A conclusion of wet-chemistry-synthesized AgGaS2 and typical AgGaS2-derived nanocrystals.
Table 2. A conclusion of wet-chemistry-synthesized AgGaS2 and typical AgGaS2-derived nanocrystals.
NanocrystalsPhasePrecursorsSize (nm)Eg (eV)Emission Range (nm)Ref
AgGaS2TetragonalAgCl, Ga, and S5.0–7.0N.A.N.A.[92]
AgGaS2TetragonalAgCl, GaCl3, and thiourea5.0N.A.446[93]
AgGaS2TetragonalAg-oleate, Ga(acac)3, S13.02.6–2.7475[82]
AgGaS2OrthorhombicAgS2CN(C2H5)2, Ga(S2CN(C2H5)2)316.3–19.12.7N.A.[94]
AgGaS2MonoclinicAgNO3, Ga(acac)3, and S3–5.6N.A.460, 650[95]
AgGaS2Orthorhombic,
rhombohedral
AgNO3, GaCl3, Na2S,2.0N.A.N.A.[96]
AgGaS2: InTetragonalAgInyGa1-y(S2CN(C2H5)2)44–5N.A.550–750[97]
AgGaS2: InTetragonalAgInS2, Ga(NO3)34.2N.A.550[98]
AgGaS2: ZnTetragonalAgNO3, Ga(acac)3, ZnSt2, and S4.2N.A.470–510[83]
AgGaS2: SeOrthorhombic, tetragonalAgNO3, GaSt3, InSt3, Se, thiourea10–181.9–2.8N.A.[99]
AgGaS2: In@ZnSTetragonalAgI, Ga(acac)3 In(Ac)3, and S5–5.32.4–2.9515–570[100]
AgGaS2: Zn@ZnSTetragonalAgI, Ga(acac)3 ZnCl2, and S5–5.32.9–3.1450–515
AgGaS2: In@ZnSTetragonalAgI, Ga(acac)3 ZnSt2, and S5.01.9–2.8460–670[101]
AgGaS2: In@ZnSTetragonalAgNO3, Ga(Ac)3, In(Ac)3, and S3.8 N.A.560–600[102]
AgGaS2: In@GaSxTetragonalAgAc, Ga(acac)3, In(acac)3, and S2.9–4.52.1–2.6500–600[103]
AgGaS2: In@GaSxTetragonalAgAc, In(Ac)3, Ga(S2CN(C2H5)2)34.32.1–2.6498–602[104]
AgGaS2@CdSeSTetragonalAgI, Ga(acac)3, and S5.1 ± 0.61.97710[105]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Xing, G.; Chen, B. AgGaS2 and Derivatives: Design, Synthesis, and Optical Properties. Nanomaterials 2025, 15, 147. https://doi.org/10.3390/nano15020147

AMA Style

Xing G, Chen B. AgGaS2 and Derivatives: Design, Synthesis, and Optical Properties. Nanomaterials. 2025; 15(2):147. https://doi.org/10.3390/nano15020147

Chicago/Turabian Style

Xing, Guansheng, and Bing Chen. 2025. "AgGaS2 and Derivatives: Design, Synthesis, and Optical Properties" Nanomaterials 15, no. 2: 147. https://doi.org/10.3390/nano15020147

APA Style

Xing, G., & Chen, B. (2025). AgGaS2 and Derivatives: Design, Synthesis, and Optical Properties. Nanomaterials, 15(2), 147. https://doi.org/10.3390/nano15020147

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop