Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

High-efficiency liquid luminescent solar concentrator based on CsPbBr3 quantum dots

Open Access Open Access

Abstract

The performance degradation is still a challenge for the development of conventional polymer luminescent solar concentrator (LSC). Liquid LSC (L-LSC) may be an alternative due to polymerization-free fabrication. Here, we have prepared a CsPbBr3 quantum dots (QDs)-based L-LSC by injecting the QDs solution into a self-assembly quartz glass mold. The as-fabricated L-LSC performance is evaluated by optical characterization and photo-electrical measurement. The external quantum efficiency of the L-LSC is up to 13.44%. After coupling the commercial solar cell, the optimal optical efficiency reaches 2.32%. These results demonstrate that L-LSC may provide a promising direction for advanced solar light harvesting technologies.

© 2022 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

Luminescent solar concentrators (LSCs) have attracted significant attention as promising solar energy conversion devices for building integrated photovoltaic (BIPV) systems due to their simple architecture and low-cost fabrication [14]. A typical LSCs is made up of luminophores and a transparent waveguide, in which luminophores dispersed into the transparent waveguide absorb solar radiation and reemit new photons at a longer wavelength. These reemitted photons are propagated to the edge of the LSCs through total internal reflection and are then absorbed by the solar cells [5,6]. So far, various types of luminophores have been investigated such as organic dyes [7,8] or colloidal quantum dots (QDs) [9,10]. In particular, all-inorganic perovskite QDs of CsPbX3 (X = Cl, Br, I) are considered as ideal luminophores [11,12] due to their narrow emission, size-dependent luminescence behavior, as well as high photoluminescence quantum yield (PLQY), and hence have found wide applications in lasers [13,14], white light-emitting diodes (WLEDs) [15,16], solar cells [17], X-ray scintillators [18], etc. More attractively, Mo et al. have synthesized a high-PLQY (80%) ZrO2-coated CsPbBr3 QDs at room temperature, realizing a high data rate of 33.5 Mbps in visible light communication [19]. It is thus expected that CsPbX3 QDs could be an ideal candidate for the fabrication of high-performance LSCs.

Besides the optical properties of the luminophores, the waveguide materials also determine the performance of the LSCs. In most of the reported cases, the luminophores are mixed with the monomer such as methyl methacrylate (MMA), and the mixture is polymerized into a bulk luminophores/polymer composite where the polymerization is driven by the initiator, forming the conventional bulk polymer LSCs [20]. However, the phase transition originating from the incorporation of luminophores with polymer tends to induce the degradation and agglomeration of luminophores, which leads to luminescence quenching, and finally diminishes the performance of LSCs [21].

To date, different methods have been reported to partially overcome these problems. One strategy involves growing a thick shell on the surface of luminophores to reduce the interaction between the luminophores and the polymer [22,23]. For instance, Meinardi et al. incorporated the CdSe/CdS QDs with giant shells (g-QDs) into polymethylmethacrylate (PMMA) to fabricate LSCs [24]. The giant shells not only isolate the core from the environment but also separate the energy between the absorption and emission spectra to increase the Stokes shift, resulting in optical efficiencies of over 10% for the g-QDs LSCs. However, a PLQY drop is found in PMMA, in which the initiator azobisisobutyronitrile (AIBN) is used to drive thermal polymerization leading to luminescence quenching in the g-QDs. Another approach is by replacing a bulk polymer LSC waveguide with a layered structure, which coats luminophores/polymer composite on a transparent substrate [25,26]. Li et al. have reported large-area thin-film LSCs (90 × 30 cm2) by depositing silicon-covered CdSe/Cd1-xZnxS QDs/polyvinylpyrrolidone (PVP) composite on transparent glass by using a doctor-blade deposition technique [27]. The reabsorption and scattering losses of the thin-film LSC are effectively “diluted” by separating the photoluminescence (PL) and waveguide region. The large-area thin-film LSCs exhibit internal quantum efficiency of over 10%. However, the refractive index mismatch between the luminophores/polymer composite and the substrate, which may lead to additional reabsorption [28], is still a problem for the large-size thin-film LSCs.

Compared to bulk polymer and thin-film LSCs, liquid LSCs (L-LSCs) in which the luminophores are dispersed in a liquid medium have some unique advantages [29,30]. For example, the PLQY of the luminophores can be perfectly preserved due to the non-phase transition. Liu et al. have reported that the L-LSCs based on Cu-doped ZnInSe QDs show a twofold optical efficiency than that of the polymer LSC [31]. Moreover, the waveguides can be recycled by replacing different luminophores, which reduces costs and improves the sustainability of L-LSCs. However, studies on L-LSCs are still scarce and their potential should be further explored. In this work, we fabricate the L-LSCs based on all-inorganic perovskite CsPbBr3 QDs by injecting the QDs solution into self-assembly quartz glass molds. A comparative experiment of the CsPbBr3 QDs in n-hexane and off-stoichiometric thiol-ene (OSTE) polymer is carried out. Compared with the PLQY of the QD solution, the PLQY of the QDs/OSTE composite decreases dramatically due to the phase transition caused by polymerization. It is thus desired that liquid LSCs can exhibit higher PL performance. In the actual test, the internal quantum efficiency and external quantum efficiency of the as-fabricated LSCs reach values of 32% and 13.44%, respectively. After coupling the solar cell, the as-fabricated LSCs show an optimal optical efficiency of 2.32% under the radiation of the standard AM 1.5 solar simulator.

2. Experimental section

2.1. Synthesis of CsPbBr3 QDs

The CsPbBr3 QDs were synthesized via a traditional hot injection method [32]. PbBr2 (0.745 g), oleylamine (OLA, 3 mL), oleic acid (OA, 1 mL), and octadecene (ODE, 24 mL) were added to a 50-mL 3-neck flask and stirred evenly followed by heating the solution at 120°C for 1 hour under an N2 atmosphere. With the complete dissolution of PbBr2, the heating temperature was raised to 180°C and then Cs-oleate (2 mL) was quickly injected into the solution. And 10 seconds later, the solution was cooled down by using an ice bath. Subsequently, ethyl acetate was added to the solution. Then the suspension was centrifuged and the supernatant was discarded. The resultant QDs were redispersed in n-hexane for further use.

2.2. Fabrication of L-LSCs based on CsPbBr3 QDs

The detailed fabrication process is described in Fig. 1. Firstly, a sheet of quartz glass with a dimension of 5 cm × 5 cm × 0.2 cm was used as a substrate and UV glue was applied to the edge of the substrate to prevent the mold from leaking. Then a quartz glass frame with an inner diameter of 4.5 cm × 4.5 cm × 0.1 cm was placed on the substrate, and a small 2 mm opening was cut out on one side of the frame to allow injection of the CsPbBr3 QDs solution. Such structure was exposed to UV light until the UV glue was completely solidified. Subsequently, another piece of quartz glass with the same dimension was attached to the top of the frame by using the same method. In this way, a quartz glass mold with a cavity was successfully fabricated. Finally, CsPbBr3 QDs solution was injected into the mold from the small hole. An L-LSC based on CsPbBr3 QDs was fabricated after the small port was sealed with UV glue. The inset shows the photo of as-fabricated LSC with a 1.5 mg/ml concentration of CsPbBr3 QDs solution.

 figure: Fig. 1.

Fig. 1. Schematic of the fabrication process of CsPbBr3 QDs-based L-LSC.

Download Full Size | PDF

2.3. Integration of commercial solar cells and L-LSCs

Four identical commercial Si solar cells (5 cm × 0.5 cm), where the open-circuit voltage (Voc) is 0.58 V, short-circuit current density (Jsc) is 32.14 mA/cm2, filling factor (FF) is 60.42%, and the power conversion efficiency (PCE) is 12.15%, were attached to each of the four edges of the synthesized L-LSCs by using UV glue. The L-LSC@Si solar cells were then irradiated with UV lamp for 10 min until solidification, in order to further measure the photo-electrical property.

2.4. Characterization

The transmission electron microscopy (TEM) characterization of CsPbBr3 QDs was acquired by using JEOL 2100F operated at 200 kV. Absorption spectra were obtained by using a double-beam Agilent Cary-5000 UV-Vis-NIR spectrophotometer with a scanning speed of 600 nm per minute. Steady-state PL spectra were operated by using the Edinburgh Fluorescence spectrometer FLS1000. The PLQY was measured with the integrating sphere and then directly calculated by the following equation,

$$PLQY = \frac{{{N_{PL}}}}{{{N_{abs}}}} = \frac{{\int {({{I_{sample}}(\lambda ) - {I_{reference}}(\lambda )} )d\lambda } }}{{\int {({{E_{sample}}(\lambda ) - {E_{reference}}(\lambda )} )d\lambda } }}, $$
where Isample and Esample are the intensity of the emission and excitation light after correction for the QD samples, respectively. Ireference and Ereference are the intensity of the emission and excitation light after correction for the reference samples (pure n-hexane solution) in a quartz cuvette, respectively. The PLQY of QDs/OSTE nanocomposite (by encapsulating the QDs in the OSTE matrix) was characterized in the same way as that of the CsPbBr3 QDs solution. The J-V curves of the samples were measured by using the Newport-Oriel solar cells energy conversion efficiency test system under the radiation of a standard AM 1.5 solar simulator (100 mW/cm2).

3. Experimental results

3.1. Structural and spectral properties of the CsPbBr3 QDs

The three-dimensional crystal structure of CsPbBr3 QDs is presented in Fig. 2(a). It shows that the Pb ion is coordinated with six bromine ions in a corner-sharing octahedral configuration, with the Cs ion located between those octahedra. The valence band of QDs mainly consists of antibonding orbitals formed by the hybridization of the 6s orbital of the Pb ion and the 4p orbital of the Br ion, while the conduction band is dominated by the antibonding orbital of the 6p orbital of the Pb ion. Based on such a special orbital character, the defect energy levels of the QDs are enclosed within the energy band, which almost has no significant effect on the radiative recombination between the band gaps, and thus enables produce a relatively high PLQY [33] . According to this, CsPbBr3 QDs are synthesized through a colloidal hot-injection method in this work, as depicted in the experimental section. The TEM image shown in Fig. 2(b) demonstrates that the as-synthesized CsPbBr3 QDs have a uniform cubic shape with high crystallinity. The lattice spacing of 0.41 nm (inset of Fig. 2(b)) corresponds to the (110) plane of the cubic-phase CsPbBr3 QDs. As shown in Fig. 2(c), the particle size histogram of as-synthesized QDs exhibits a Gauss distribution featuring an average size of 7.56 ± 1.23 nm. Figure 2(d) shows the absorption and PL spectra of the CsPbBr3 QDs. A wide absorption range from 300 nm to 510 nm and a narrow green emission band peaked at around 517 nm with a full width at half maximum of 20 nm are observed. The PL emission of the CsPbBr3 QDs can be perfectly matched with the external quantum efficiency (EQE) spectrum of the Si solar cell, proving that the CsPbBr3 QD is an ideal candidate for the luminescence centers of LSC.

 figure: Fig. 2.

Fig. 2. (a) The three-dimensional crystal structure, (b) TEM image (Inset is HTEM of QDs), (c) Particle size histogram, and (d) Absorption and PL spectra of as-synthesized CsPbBr3 QDs.

Download Full Size | PDF

3.2. Comparative study between the QDs solution and the QDs/polymer composite

The effect of phase transition from liquid to solid on the optical properties of the CsPbBr3 QDs is investigated. As shown in Fig. 3(a), the position and shape of the emission peaks of CsPbBr3 QDs distributed in n-hexane (QDs solution), OSTE without photo-initiator as well as in OSTE after polymerization with photo-initiator are almost unchanged, which indicates that the polymerization process does not result in the formation of a considerable density of surface defects that usually introduce additional non-radiative channels [31]. Compared with that of the pure QDs solution, however, the PL intensity of QDs/OSTE composites is obviously decreased. The digital photo in the inset of Fig. 3(a) also reflects the QDs/OSTE composites emit much weaker green light than the pure QDs under UV-light irradiation (365 nm). The reduction in the PL intensity for the QDs/OSTE composites can be attributed to the poor compatibility between QDs and OSTE, which leads to the degradation of the QDs. With some initiators added into the QDs/OSTE composites during the polymerization process, the chemical attack of the CsPbBr3 QDs by the initiator radicals inevitably leads to the luminescence quenching. Meanwhile, the heat released from the polymerization process also reduces the PL intensity of the CsPbBr3 QDs due to the thermal instability caused by the low formation energy of such perovskite lattices [34]. Another detrimental effect is associated with the aggregation of the QDs during polymerization. The formed large-size QD clusters are easily coupled to the small-size QDs, leading to dot-to-dot energy transfer and weakening the luminescence consequently [27]. Figures 3(b), (c) show that the PL decay curve of QDs in initiator-added OSTE exhibits a shorter lifetime (3.59 ns) than that in n-hexane (8.79 ns), and the PLQY of polymerized QDs/OSTE composite (9%) is 10 times lower than that of QDs solution (89%) under the 405 nm excitation. All of these observations indicate that fabricating liquid state LSC without PL loss seems to be a more efficient strategy to obtain high-performance LSC.

 figure: Fig. 3.

Fig. 3. (a) PL spectra of CsPbBr3 QDs in n-hexane (black line, No.1-Liquid state), in OSTE without photo-initiator (red line, No.2-Unpolymerized), and in OSTE after polymerization with photo-initiator (blue line, No.3-Solid state). (b) PL decay curves of CsPbBr3 QDs in liquid state (black dot) and solid state (red dot). (c) PL spectra of CsPbBr3 QDs in liquid state (black line) and solid-state (red line), as well as the background (blue line) collected at an integrating sphere.

Download Full Size | PDF

3.3. Optical characterization of the L-LSCs

Herein, the L-LSCs have been fabricated by injecting different concentrations of CsPbBr3 QDs into self-assembling quartz molds, and the detail is described in the experimental section. To evaluate the potential of the resultant L-LSCs, we first investigated the total PL quantum yield (ηPL,LSC) of L-LSCs. The specific calculation process of ηPL,LSC is described in Fig. S1 (Supplement 1). Figure 4(a) shows the ηPL,LSC is increased from 64% to 96% with the QDs concentration increasing from 0.5 mg/ml to 1.5 mg/ml. However, by further adding QDs concentration, the ηPL,LSC suffers a decrease, which is contributed to the energy loss originating from the non-radiative relaxation due to the concentration quenching.

 figure: Fig. 4.

Fig. 4. (a) The total PL quantum yield, (b) The edge-emission efficiency, (c) The internal quantum efficiency, (d) The absorption spectra, and (e) The ηabs and ηext of the L-LSCs with different concentrations of CsPbBr3 QDs.

Download Full Size | PDF

Figure 4(b) shows the edge-emitting efficiency (ηedge) with different concentrations of CsPbBr3 QDs. The calculation process of ηedge is provided in Fig. S2 (Supplement 1). The optimal ηedge of the L-LSCs is 34%, corresponding to the CsPbBr3 QDs concentration of 1.5 mg/ml. A further increase in QDs concentration leads to a decrease in ηedge. This decreasing ηedge may originate from the ‘randomization’ of photon propagation direction during each re-absorption, which results in additional losses via re-emitted photons into the escape cone [1,19]. According to ηPL,LSC and ηedge, the internal quantum efficiency (ηint) of the L-LSC can be obtained by calculating as,

$${\eta _{{\mathop{\rm int}} }} = {\eta _{PL,LSC}} \times {\eta _{edge}}. $$

As shown in Fig. 4(c), it can be seen that the highest ηint of CsPbBr3 QDs-based L-LSCs is 32%, which is relatively higher than that of the previously reported polymer LSCs (Table 1).

Tables Icon

Table 1. The comparison of ηint and ηext for LSCs based on different luminophores.

The external quantum efficiency (ηext) is another important parameter to assess the performance of the LSCs, involving the response to the full spectrum of sunlight. ηext of LSCs is further calculated using the Eq. (3),

$${\eta _{ext}} = {\eta _{abs}} \times {\eta _{{\mathop{\rm int}} }}, $$
where ηabs is the absorption efficiency of LSC calculated as,
$${\eta _{abs}} = (1 - R)\frac{{\int_0^\infty {{S_{in}}(\lambda )(1 - {e^{ - \alpha (\lambda )d}})d\lambda } }}{{\int_0^\infty {{S_{in}}(\lambda )d\lambda } }}, $$
$$R = \frac{{{{({{n_2} - {n_1}} )}^2}}}{{{{({{n_2} + {n_1}} )}^2}}}, $$
where Sin(λ) is the solar radiation intensity obtained using a solar simulator. α(λ) is the absorption coefficient of the LSCs which is calculated based on the absorption spectra in Fig. 4(d). R is reflectivity, calculated from Eq. (5) to be 0.04, where the refractive index nquartz = 1.5 and nair = 1. d is the thickness of LSCs (0.5 cm). Figure 4(e) shows the ηabs and ηext as a function of CsPbBr3 QDs concentration, respectively. Since the absorption of LSCs is proportional to the QDs concentration (Fig. 4(d)), the ηabs increase gradually with increasing the QDs concentration. However, the decline of ηint appearing at the CsPbBr3 QDs concentration above 1.5 mg/ml leads to a reduction of ηext consequently. The optimal ηext for the as-fabricated LSCs is 13.44%. A comparison of ηext for different types of LSCs is listed in Table 1. It can be seen that the prepared L-LSCs exhibit a relatively high ηext compared to the reported polymer LSCs, implying that the L-LSCs based on CsPbBr3 QDs may be an alternative for solar light harvesting applications.

The stability of the L-LSC with 1.5 mg/ml QDs concentration under solar irradiation has been investigated (Fig. S3, Supplement 1). It shows the PL intensity of the L-LSC dropped to 81% of the initial intensity after 15 days of solar irradiation, while that of the QDs solution were down to 60%, which indicates the sealed glass mold can protect the QDs from water and oxygen to some extent.

3.4. Photo-electrical measurement of the L-LSCs

The photo-electrical properties of the L-LSCs are measured to further illustrate the light collection performance, and the test method is according to our previous work [39]. Figure 5(a) shows the J-V curves of the as-fabricated LSCs with different concentrations of CsPbBr3 QDs. The short-circuit current density for the as-fabricated LSCs is initially enhanced with the concentration of the QDs increasing from 0.5 mg/ml to 1.5 mg/ml, while further increasing the QD concentration gives rise to a decline. The maximum of the short-circuit current density is 1.87 mA/cm2 corresponding to the CsPbBr3 QDs concentration of 1.5 mg/ml. The optical efficiency (ηopt) of the LSCs is calculated as,

$${\eta _{opt}} = \frac{{{I_{sc - LSC}}}}{{{I_{sc}} \times G}}, $$
where Isc-LSC and Isc are the short circuit currents from the solar cell attached and not attached to the LSCs under direct solar radiation, respectively. G is the geometric gain factor defined as the ratio of the top area of LSC to the edge area of LSC, which is calculated to be 2.5 for the as-fabricated L-LSCs. Figure 5(b) shows the ηopt as a function of CsPbBr3 QDs concentration. Based on the ηopt, the EQE spectra of the L-LSC with different concentrations of CsPbBr3 QDs are also studied and the corresponding results are given in Fig. S4 (Supplement 1). The optimal ηopt for the as-fabricated L-LSCs with the CsPbBr3 QDs concentration of 1.5 mg/ml is 2.32%. The relatively low ηopt may be contributed to the coupling loss at the interface between the LSC and the solar cell. Thus, there is still room for improving the efficiency of L-LSCs by optimizing the fabrication process in future work. Notably, the self-assembled quartz mold of liquid LSCs can be easily recycled by injecting different luminophores, reducing the cost of the fabrication of LSCs.

 figure: Fig. 5.

Fig. 5. (a) The J-V curves and (b) the ηopt of the L-LSCs with different concentrations of CsPbBr3 QDs.

Download Full Size | PDF

4. Conclusion

The high-efficiency liquid LSCs based on perovskite CsPbBr3 QDs have been realized. Compared to pure QDs solution, the QDs dispersed in polymer suffer from a serious luminescence quenching due to the phase transition. The liquid LSCs are thus fabricated by injecting the QDs solution into the self-assembled quartz glass molds. Benefiting from this polymerization-free fabrication, the resultant LSCs exhibit high internal quantum efficiency (32%) and external quantum efficiency (13.44%) at the CsPbBr3 QDs concentration of 1.5 mg/ml. Such values are higher than that of the reported polymer LSCs. Under the radiation of a standard AM 1.5 solar simulator, the optical efficiency of the LSCs is up to 2.32%. We thus anticipate that this high-efficient and low-cost liquid LSC should be a promising strategy for the future application of LSC.

Funding

National Natural Science Foundation of China (51702172, 61974078); Natural Science Foundation of Zhejiang Province (LY20A040002, LY21F040002); Fundamental Research Funds for the Provincial Universities of Zhejiang (SJLY2021012); Key Research and Development Program of Jiangsu Province (BE2021082); Natural Science Foundation of Ningbo (2021J059).

Disclosures

The authors declare that there are no conflicts of interest related to this article.

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

Supplemental document

See Supplement 1 for supporting content.

References

1. I. Papakonstantinou, M. Portnoi, and M. G. Debije, “The hidden potential of luminescent solar concentrators,” Adv. Energy Mater. 11(3), 2002883 (2021). [CrossRef]  

2. M. Raeisossadati, N. R. Moheimani, and D. Parlevliet, “Luminescent solar concentrator panels for increasing the efficiency of mass microalgal production,” Renewable Sustainable Energy Rev. 101, 47–59 (2019). [CrossRef]  

3. G. H. Timmermans, B. W. Saes, and M. G. Debije, “Dual-responsive “smart” window and visually attractive coating based on a diarylethene photochromic dye,” Appl. Opt. 58(36), 9823–9828 (2019). [CrossRef]  

4. B. Zhang, G. Lyu, E. A. Kelly, and R. C. Evans, “Förster resonance energy transfer in luminescent solar concentrators,” Adv. Sci. 9(23), 2201160 (2022). [CrossRef]  

5. K. B. Cai, H. Y. Huang, M. L. Hsieh, P. W. Chen, S. E. Chiang, S. H. Chang, J. L. Shen, W.-R. Liu, and C. T. Yuan, “Two-dimensional self-assembly of boric acid-functionalized graphene quantum dots: tunable and superior optical properties for efficient eco-friendly luminescent solar concentrators,” ACS Nano 16(3), 3994–4003 (2022). [CrossRef]  

6. I. Sychugov, “Analytical description of a luminescent solar concentrator,” Optica 6(8), 1046–1049 (2019). [CrossRef]  

7. C. Liu and B. Li, “Multiple dyes containing luminescent solar concentrators with enhanced absorption and efficiency,” J. Opt. 17(2), 025901 (2015). [CrossRef]  

8. M. Peng, B. Dong, X. Cai, W. Wang, X. Jiang, Y. Wang, Y. Yang, and D. Zou, “Organic dye-sensitized photovoltaic fibers,” Sol. Energy 150(1), 161–165 (2017). [CrossRef]  

9. A. Kim, A. Hosseinmardi, P. K. Annamalai, P. Kumar, and R. Patel, “Review on colloidal quantum dots luminescent solar concentrators,” ChemistrySelect 6(20), 4948–4967 (2021). [CrossRef]  

10. X. Wang, Y. Zhang, J. Li, G. Liu, M. Gao, S. Ren, B. Liu, L. Zhang, G. Han, and J. Yu, “Platinum cluster/carbon quantum dots derived graphene heterostructured carbon nanofibers for efficient and durable solar-driven electrochemical hydrogen evolution,” Small Methods 6(4), 2101470 (2022). [CrossRef]  

11. S. Li, M. Yuan, W. Zhuang, X. Zhao, S. Tie, J. Xiang, and S. Lan, “Optically-controlled quantum size effect in a hybrid nanocavity composed of a perovskite nanoparticle and a thin gold film,” Laser Photonics Rev. 15(3), 2000480 (2021). [CrossRef]  

12. S. Zhao, Q. Mo, B. Wang, W. Cai, R. Li, and Z. Zang, “Inorganic halide perovskites for lighting and visible light communication,” Photonics Res. 10(4), 1039–1062 (2022). [CrossRef]  

13. S. Lin, H. Lin, G. Chen, B. Wang, X. Yue, Q. Huang, J. Xu, Y. Cheng, and Y. Wang, “Stable CsPbBr3-glass nanocomposite for low-étendue wide-color-gamut laser-driven projection display,” Laser Photonics Rev. 15(7), 2100044 (2021). [CrossRef]  

14. C. Li, Z. Zang, C. Han, Z. Hu, X. Tang, J. Du, Y. Leng, and K. Sun, “Highly compact CsPbBr3 perovskite thin films decorated by ZnO nanoparticles for enhanced random lasing,” Nano Energy 40, 195–202 (2017). [CrossRef]  

15. K. C. Shen, J. K. Wang, Y. Shen, Y. Q. Li, M. L. Guo, Z. H. Su, L. Y. Lu, X. Y. Cai, L. Chen, and F. Song, “Unraveling the role of crystallization dynamics on luminescence characteristics of perovskite light-emitting diodes,” Laser Photonics Rev. 15(7), 2100023 (2021). [CrossRef]  

16. D. Yan, S. Zhao, Y. Zhang, H. Wang, and Z. Zang, “Highly efficient emission and high-CRI warm white light-emitting diodes from ligand-modified CsPbBr3 quantum dots,” Opto-Electron. Adv. 5(1), 200075 (2022). [CrossRef]  

17. J. P. Correa-Baena, M. Saliba, T. Buonassisi, M. Grätzel, A. Abate, W. Tress, and A. Hagfeldt, “Promises and challenges of perovskite solar cells,” Science 358(6364), 739–744 (2017). [CrossRef]  

18. B. Wang, J. Peng, X. Yang, W. Cai, H. Xiao, S. Zhao, Q. Lin, and Z. Zang, “Template assembled large-size CsPbBr3 nanocomposite films toward flexible, stable, and high-performance X-ray scintillators,” Laser Photonics Rev. 16(7), 2100736 (2022). [CrossRef]  

19. Q. Mo, C. Chen, W. Cai, S. Zhao, D. Yan, and Z. Zang, “Room temperature synthesis of stable zirconia-coated CsPbBr3 nanocrystals for white light-emitting diodes and visible light communication,” Laser Photonics Rev. 15(10), 2100278 (2021). [CrossRef]  

20. Y. Li, X. Zhang, Y. Zhang, R. Dong, and C. K. Luscombe, “Review on the role of polymers in luminescent solar concentrators,” J. Polym. Sci., Part A: Polym. Chem. 57(3), 201–215 (2019). [CrossRef]  

21. J. Bomm, A. Büchtemann, A. Fiore, L. Manna, J. H. Nelson, D. Hill, and W. G. van Sark, “Fabrication and spectroscopic studies on highly luminescent CdSe/CdS nanorod polymer composites,” Beilstein J. Nanotechnol. 1, 94–100 (2010). [CrossRef]  

22. H. Bahmani Jalali, S. Sadeghi, I. Baylam, M. Han, C. W. Ow-Yang, A. Sennaroglu, and S. Nizamoglu, “Exciton recycling via InP quantum dot funnels for luminescent solar concentrators,” Nano Res. 14(5), 1488–1494 (2021). [CrossRef]  

23. S. Hase, Y. Iso, and T. Isobe, “Bandgap-tuned fluorescent CuGaS 2/ZnS core/shell quantum dots for photovoltaic applications,” J. Mater. Chem. C 10(9), 3523–3530 (2022). [CrossRef]  

24. F. Meinardi, A. Colombo, K. A. Velizhanin, R. Simonutti, M. Lorenzon, L. Beverina, R. Viswanatha, V. I. Klimov, and S. Brovelli, “Large-area luminescent solar concentrators based on ‘Stokes-shift-engineered’ nanocrystals in a mass-polymerized PMMA matrix,” Nat. Photonics 8(5), 392–399 (2014). [CrossRef]  

25. J. Wang, Y. Yuan, J. Schneider, W. Zhou, H. Zhu, T. Cai, and O. Chen, “Quantum dot-based luminescent solar concentrators fabricated through the ultrasonic spray-coating method,” ACS Appl. Mater. Interfaces 14(36), 41013–41021 (2022). [CrossRef]  

26. Z. Zheng, Y. Zhang, X. Cao, G. Gu, Y. Tian, and X. Zhang, “Modeling and comparison of bulk and thin-film luminescent solar concentrators based on colloidal perovskite quantum dots,” Opt. Lett. 47(17), 4367–4370 (2022). [CrossRef]  

27. H. Li, K. Wu, J. Lim, H.-J. Song, and V. I. Klimov, “Doctor-blade deposition of quantum dots onto standard window glass for low-loss large-area luminescent solar concentrators,” Nat. Energy 1(12), 16157 (2016). [CrossRef]  

28. J. Roncali, “Luminescent solar collectors: quo vadis ?” Adv. Energy Mater. 10(36), 2001907 (2020). [CrossRef]  

29. S. Sadeghi, R. Melikov, H. Bahmani Jalali, O. Karatum, S. B. Srivastava, D. Conkar, E. N. Firat-Karalar, and S. Nizamoglu, “Ecofriendly and efficient luminescent solar concentrators based on fluorescent proteins,” ACS Appl. Mater. Interfaces 11(9), 8710–8716 (2019). [CrossRef]  

30. C. Tummeltshammer, M. Portnoi, S. A. Mitchell, A.-T. Lee, A. J. Kenyon, A. B. Tabor, and I. Papakonstantinou, “On the ability of Förster resonance energy transfer to enhance luminescent solar concentrator efficiency,” Nano Energy 32, 263–270 (2017). [CrossRef]  

31. X. Liu, B. Luo, J. Liu, D. Jing, D. Benetti, and F. Rosei, “Eco-friendly quantum dots for liquid luminescent solar concentrators,” J. Mater. Chem. A 8(4), 1787–1798 (2020). [CrossRef]  

32. G. Nedelcu, L. Protesescu, S. Yakunin, M. I. Bodnarchuk, M. J. Grotevent, and M. V. Kovalenko, “Fast anion-exchange in highly luminescent nanocrystals of cesium lead halide perovskites (CsPbX3, X = Cl, Br, I),” Nano Lett. 15(8), 5635–5640 (2015). [CrossRef]  

33. Q. A. Akkerman, G. Rainò, M. V. Kovalenko, and L. Manna, “Genesis, challenges and opportunities for colloidal lead halide perovskite nanocrystals,” Nat. Mater. 17(5), 394–405 (2018). [CrossRef]  

34. S. Zou, Y. Liu, J. Li, C. Liu, R. Feng, F. Jiang, Y. Li, J. Song, H. Zeng, and M. Hong, “Stabilizing cesium lead halide perovskite lattice through Mn (II) substitution for air-stable light-emitting diodes,” J. Am. Chem. Soc. 139(33), 11443–11450 (2017). [CrossRef]  

35. M. J. Talite, H.-Y. Huang, Y.-H. Wu, P. G. Sena, K.-B. Cai, T.-N. Lin, J.-L. Shen, W.-C. Chou, and C.-T. Yuan, “Greener luminescent solar concentrators with high loading contents based on in situ cross-linked carbon nanodots for enhancing solar energy harvesting and resisting concentration-induced quenching,” ACS Appl. Mater. Interfaces 10(40), 34184–34192 (2018). [CrossRef]  

36. X. Luo, T. Ding, X. Liu, Y. Liu, and K. Wu, “Quantum-cutting luminescent solar concentrators using ytterbium-doped perovskite nanocrystals,” Nano Lett. 19(1), 338–341 (2019). [CrossRef]  

37. S. Sadeghi, H. B. Jalali, S. B. Srivastava, R. Melikov, I. Baylam, A. Sennaroglu, and S. Nizamoglu, “High-performance, large-area, and ecofriendly luminescent solar concentrators using copper-doped InP quantum dots,” iScience 23(7), 101272 (2020). [CrossRef]  

38. L. Zdražil, S. Kalytchuk, K. Holá, M. Petr, O. Zmeškal, Š Kment, A. L. Rogach, and R. Zbořil, “A carbon dot-based tandem luminescent solar concentrator,” Nanoscale 12(12), 6664–6672 (2020). [CrossRef]  

39. G. Gu, X. Zhang, Y. Guo, Z. Zheng, X. Cao, Y. Zhang, R. Huang, and X. Zhang, “Optical characterization and photo-electrical measurement of luminescent solar concentrators based on perovskite quantum dots integrated into the thiol-ene polymer,” Appl. Phys. Lett. 119(1), 011905 (2021). [CrossRef]  

Supplementary Material (1)

NameDescription
Supplement 1       Experimental characterization

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (5)

Fig. 1.
Fig. 1. Schematic of the fabrication process of CsPbBr3 QDs-based L-LSC.
Fig. 2.
Fig. 2. (a) The three-dimensional crystal structure, (b) TEM image (Inset is HTEM of QDs), (c) Particle size histogram, and (d) Absorption and PL spectra of as-synthesized CsPbBr3 QDs.
Fig. 3.
Fig. 3. (a) PL spectra of CsPbBr3 QDs in n-hexane (black line, No.1-Liquid state), in OSTE without photo-initiator (red line, No.2-Unpolymerized), and in OSTE after polymerization with photo-initiator (blue line, No.3-Solid state). (b) PL decay curves of CsPbBr3 QDs in liquid state (black dot) and solid state (red dot). (c) PL spectra of CsPbBr3 QDs in liquid state (black line) and solid-state (red line), as well as the background (blue line) collected at an integrating sphere.
Fig. 4.
Fig. 4. (a) The total PL quantum yield, (b) The edge-emission efficiency, (c) The internal quantum efficiency, (d) The absorption spectra, and (e) The ηabs and ηext of the L-LSCs with different concentrations of CsPbBr3 QDs.
Fig. 5.
Fig. 5. (a) The J-V curves and (b) the ηopt of the L-LSCs with different concentrations of CsPbBr3 QDs.

Tables (1)

Tables Icon

Table 1. The comparison of ηint and ηext for LSCs based on different luminophores.

Equations (6)

Equations on this page are rendered with MathJax. Learn more.

P L Q Y = N P L N a b s = ( I s a m p l e ( λ ) I r e f e r e n c e ( λ ) ) d λ ( E s a m p l e ( λ ) E r e f e r e n c e ( λ ) ) d λ ,
η int = η P L , L S C × η e d g e .
η e x t = η a b s × η int ,
η a b s = ( 1 R ) 0 S i n ( λ ) ( 1 e α ( λ ) d ) d λ 0 S i n ( λ ) d λ ,
R = ( n 2 n 1 ) 2 ( n 2 + n 1 ) 2 ,
η o p t = I s c L S C I s c × G ,
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.