ArticlePDF Available

Efficient and Stable Thin‐Film Luminescent Solar Concentrators Enabled by Near‐Infrared Emission Perovskite Nanocrystals

Authors:

Abstract and Figures

TPP für LSCs: Eine Nachbehandlung mit Triphenylphosphin (TPP) wurde für die Herstellung von hochstabilen, großflächigen lumineszierenden Solarkonzentratoren (LSCs) aus CsPbI3‐Polymer‐Kompositfilmen entwickelt. Die LSCs haben eine hohe optische Konversionseffizienz und sind hoch emittierend (absoluter PLQY ca. 100 %). Abstract A novel triphenylphosphine (TPP) treatment strategy was developed to prepare the near‐infrared emission CsPbI3 nanocrystal (NC)‐polymer composite thin‐film luminescent solar concentrators (LSCs) featuring high absolute photoluminescence quantum yield (PLQY), low reabsorption, and high stability. The PL emission of the LSCs is centered at about 700 nm with 99.4±0.4 % PLQY and narrow full width at half maximum (FWHM) of 75 meV (30 nm). Compared with LSCs prepared with classic CsPbI3 NCs, the stability of the LSCs after TPP treatments has been greatly improved, even after long‐term (30 days) immersion in water and strong mercury‐lamp irradiation (50 mW cm⁻²). Owing to the presence of lone‐pair electrons on the phosphorus atom, TPP is also used as a photoinitiator, with higher efficiency than other common photoinitiators. Large‐area (ca. 75 cm²) infrared LSCs were achieved with a high optical conversion efficiency of 3.1 % at a geometric factor of 10.
Content may be subject to copyright.
Perovskites
Efficient and Stable Thin-Film Luminescent Solar Concentrators
Enabled by Near-Infrared Emission Perovskite Nanocrystals
Jiajing Wu, Jianyu Tong, Yuan Gao, Aifei Wang, Tao Zhang, Hairen Tan, Shuming Nie, and
Zhengtao Deng*
Abstract: A novel triphenylphosphine (TPP) treatment strat-
egy was developed to prepare the near-infrared emission
CsPbI3nanocrystal (NC)-polymer composite thin-film lumi-
nescent solar concentrators (LSCs) featuring high absolute
photoluminescence quantum yield (PLQY), low reabsorption,
and high stability. The PL emission of the LSCs is centered at
about 700 nm with 99.40.4% PLQY and narrow full width
at half maximum (FWHM) of 75 meV (30 nm). Compared
with LSCs prepared with classic CsPbI3NCs, the stability of
the LSCs after TPP treatments has been greatly improved, even
after long-term (30 days) immersion in water and strong
mercury-lamp irradiation (50 mWcm2). Owing to the pres-
ence of lone-pair electrons on the phosphorus atom, TPP is
also used as a photoinitiator, with higher efficiency than other
common photoinitiators. Large-area (ca. 75 cm2) infrared
LSCs were achieved with a high optical conversion efficiency
of 3.1% at a geometric factor of 10.
Colloidal nanocrystals (NCs) are of great interest for
a variety of high-performance optical and optoelectronic
applications, such as bioimaging,[1] solar cells,[2] light-emitting
diodes,[3] photodetectors,[4] and luminescent solar concentra-
tors (LSCs).[5,12] In particular, compared with traditional
organic dyes, NCs are being widely used to develop large-area
LSCs owing to their advantages of high PLQY, tunable
absorption/emission spectra, and good stability.[5a,d,6] At
present, these NCs mainly include C-dots,[7] Si,[8] CdSe/
CdS,[12] CdSe/CdZnS,[5d] CdSe/CdPbS,[5c] Mn2+-doped ZnSe/
ZnS,[9] CuInS2/ZnS,[5e] CuInSSe/ZnS,[10] CuInSeS/ZnS,[11] PbS/
CdS.[5b] Among them, the optical conversion efficiency of
2.86% has been reported based on Si NCs.[8] Optical
conversion efficiency of LSC-based giant CdSe/CdS NCs
could reach 48% under single-wavelength excitation.[12]
However, it is a major challenge to fabricate NC-based
LSCs with high PLQY and long-term stability under outdoor
conditions.
In recent years, lead halide perovskites (APbX3A=MA+,
FA+or Cs+;X=I,Br
,Cl
) have attracted extensive
attention owing to their excellent performance and low
processing cost.[13] An attractive feature of these materials is
a tunable band gap between 1.48 eV and 2.3 eV, which is
required for LSCs.[13b,14] Recently, some researchers have
begun to prepare LSCs from lead halide perovskite.[14,15] For
example, Meinardi et al. demonstrate Mn-doped CsPbCl3
NCs as reabsorption-free emitters for large-area LSCs.[16]
Wei et al. fabricated low-loss large-area LSCs based on
green emission layered perovskite nanoplatelets, achieving
optical quantum efficiency of 26 % and external optical
quantum efficiency of 0.87%.[17] Our group prepared
FAPbBr3NC–polymer slabs with high PLQY ( 92 5%)
and stable green emission for LSCs.[18] With a direct band gap
of 1.73 eV, CsPbI3is very suitable for LSCs.[19] However,
traditional CsPbI3NCs have serious phase-transition prob-
lems under ambient conditions, which limit their practical
applications. To resolve this phase instability problem, many
strategies have been proposed, such as doping, surface ligand
engineering, and providing protective shells.[19,20] Several
group have reported that trioctylphosphine could enhanced
stabilization of CsPbI3quantum dots in solution.[19] Recently,
Bi et al. demonstrated a surface ligand modification of CsPbI3
by using a 2-aminoethanethiol (AET) treatment to improve
the stability of NCs both in solution and as films. However,
AET-CsPbI3NCs dispersed in solution have relatively low
PLQY of 51% with broad FWHM of 41 nm; the AET-CsPbI3
NC-films have low PLQY of only 33 %.[21] Equally important,
the long-term stability of CsPbI3NC–polymer composite
films remains a major challenge. Therefore, the development
of a more effective method for preparing CsPbI3NC–polymer
composite films is of great significance for LSCs.
In this study, for the first time, we prepared stable and
bright near-infrared emissive thin-film LSCs by UV-polymer-
ization of acrylic resin containing TPP treated CsPbI3NCs.
Our TPP treatment strategy has several advantages: 1) the
PLQY of the CsPbI3NC–polymer composite films increased
significantly from 52 % to 99.4 0.4 %; 2) the stability of the
films has been greatly improved, even after long-term
(30 day) immersion in water and strong mercury lamp
irradiation (50 mW cm2); 3) owing to the presence of lone-
pair electrons on the phosphorus atom,[22] TPP is also used as
a photoinitiator, which is more efficient than other common
initiators. Therefore, we used the TPP-treated CsPbI3NCs to
produce near-infrared LSCs devices, with an optical conver-
sion efficiency of 3.1% at the geometric factor of 10 and good
[*] J. Wu, Dr. J. Tong, Dr. Y. Gao, Dr. A. Wang, Prof. T. Zhang, Prof. H. Tan,
Prof. Z. Deng
College of Engineering and Applied Sciences, Nanjing National
Laboratory of Microstructures, Nanjing University
Nanjing, Jiangsu, 210023 (P. R. China)
E-mail: dengz@nju.edu.cn
Prof. S. Nie
Departments of Bioengineering, Chemistry, Electrical and Computer
Engineering, and Materials Science and Engineering
University of Illinois at Urbana-Champaign
Urbana, IL 61801 (USA)
Supporting information and the ORCID identification number(s) for
the author(s) of this article can be found under:
https://doi.org/10.1002/anie.201911638.
A
ngewandte
Chemi
e
Communications
How to cite:
International Edition: doi.org/10.1002/anie.201911638
German Editon: doi.org/10.1002/ange.201911638
7738 2020 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2020,59, 7738 –7742
environmental stability. In the absence of encapsulation, the
light efficiency of the LSCs can remain 93.6% after two
months of storage in the atmosphere. Therefore, we believe
that the TPP post-treatment method will be a promising
strategy for designing efficient and stable thin-film LSCs.
Near-infrared emission CsPbI3perovskite NCs were
prepared using an improved hot-injection procedure (see
the Supporting Information for details). Figure 1 a,b shows
photographs of the fresh untreated CsPbI3NCs under room
light and UV light, which is bright red when excited by 365 nm
UV light. Transmission electron microscopy (TEM) and high-
resolution TEM (HRTEM) images (Figure 1 c,d) show the
cubic shape of the as-synthesized NCs, with an average edge
length of 12.6 nm (Supporting Information, Figure S1). Fig-
ure 1e shows the powder X-ray diffraction (XRD) pattern of
CsPbI3NCs, which is a cubic-phase structure.
Figure 1f shows the typical UV/Vis absorption and PL
emission spectra of CsPbI3NCs in the toluene solution. The
UV/Vis absorption spectrum showed that there was an
absorption tail between 700 nm and 800 nm, indicating the
presence of NCs aggregates in the toluene solution. The PL
peak is located at 695 nm, and the full-width at half-maximum
(FWHM) is 30 nm.
To apply these colloidal NCs to high-performance LSCs, it
is often necessary to process them into highly transparent thin
films on the substrate. Here, we prepared CsPbI3NC–
polymer films by photopolymerization in the atmosphere
for LSCs. Using the ratio of the emission spectrum of the
TPP-CsPbI3film to the blank spectrum (Figure 2 a), the
absolute PLQY of a typical TPP-treated NC–polymer film
was calculated, and near-unity PLQY was obtained. As shown
in the Supporting Information, Figure S2, the PLQY meas-
urements were repeated 16 times, and the average value was
99.4 0.4%. Figure 2b shows the non-normalized absorption
and emission spectra of NC–polymer films. Compared to the
CsPbI3NCs in solution, the absorption edge of the NC–
polymer films showed the blue-shift without obvious absorp-
tion tails, and the PL peak position is slightly red-shifted to
700 nm with narrow FWHM of 75 meV (that is, 30 nm).
Furthermore, for comparison, we prepared CsPbI3NC–
polymer films without TPP or zinc iodide precursor in the
UV gel. Their UV/Vis absorption and PL emission spectra
were shown in Figure 2b. Compared with untreated NC
polymer films, the PL peak position of TPP treated NC–
polymer films show a slight blue-shift. It is worth noting that
the overlap of the emission and absorption of the TPP treated
NC–polymer films is smaller than that of untreated NC
polymer films (Supporting Information, Figure S4), indicating
that TPP can cause a decrease in reabsorption. The possible
reason is that, on the one hand, because TPP has a large steric
hindrance, it can improve the dispersibility of perovskite NCs
in the polymer matrix and inhibit their agglomeration. On the
other hand, TPP has strong coordination and deoxidation
ability, which can effectively protect NCs and prevent NCs
from being decomposed and regrowth during photopolyme-
rization. These will reduce the absorption tail and thus reduce
reabsorption.
To verify the stability of the TPP treated NC–polymer
composite films, we tracked the PLQYs of these films over
time in different environments. The TPP treated films show
excellent air stability and water stability (Supporting Infor-
mation, Figures S5,S6). Furthermore, photostability is an
important aspect of LSCs applications. We placed the films
in a photochemical reactor under 500 W mercury lamp
(power density 50 mWcm2, acceleration coefficient 10 at
365 nm; Supporting Information, Figure S7). As shown in
Figure 2c, the PLQY of untreated films and ZnI2-treated
Figure 1. a),b) Photographs of the as-prepared CsPbI3NCs in a three-
necked flask under room light (a) and UV light (b). c) TEM and
d) HRTEM images of CsPbI3NCs. e) The XRD pattern of CsPbI3NCs ;
inset: corresponding crystal diagram. f) UV/Vis absorption (black line)
and PL emission spectra (red line) of CsPbI3NCs; inset: photograph
of the NCs in toluene under UV light, which emitted strong red
fluorescence under UV light.
Figure 2. a) The PL emission spectra of blank (without sample, black
line) and TPP-CsPbI3(red line) used for measuring the PLQYs. Inset:
photograph of a TPP-CsPbI3NCs film in a mounting boat. b),c) UV/
Vis absorption spectra, PL emission spectra, and photostability of the
untreated, the ZnI2treated, and the TPP treated NC–polymer compo-
site films. Inset in (c): pictures of the untreated NCs films (1), ZnI2
treated films (2), and TPP treated NCs films (3) at illumination time of
0 h and 100 h.
A
ngewandte
Chemi
e
Communications
7739Angew. Chem. Int. Ed. 2020,59, 7738 –7742 2020 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.angewandte.org
films rapidly decreased to 6.4% and 10.3 %, respectively.
However, the TPP treated films still account for 88% of
PLQY under standard outdoor conditions (PLQY drops to
90% of the initial value), which is converted into 1000 h T90
under continuous light conditions. These results indicate that
TPP greatly improves the stability of the CsPbI3NC–polymer
composite films.
Interestingly, apart from being a strong ligand, TPP can
also act as a photoinitiator because there are lone pairs of
electrons on the phosphorus atoms. For comparison, various
NC–polymer composite films were fabricated using other
common initiators, such as TPO, DMPA, HPK, and HMPO.
Their structural formulas are shown in Figure 3a. Compared
to NC–polymer films treated with other initiators, the overlap
between non-normalized absorption and emission spectra of
TPP treated NC–polymer composite films is significantly
reduced (Figure 3b; Supporting Information, Figure S8). To
evaluate the photostability of these films, we tracked the
optical properties of the illumination duration (Supporting
Information, Figures S9–S11). The absorbance values of the
films with TPO, DMPA, HPK, or HMPO as initiators
changed significantly at 685 nm, while the absorbance of
TPP treated films did not change much (Figure 3c). The
PLQYs of the films treated with HMPO decreased signifi-
cantly to 25.8% after photopolymerization. With TPO,
DMPA, and HPK as initiators, PLQYs decreased to 26%,
3.5%, and 7.7% after 40 h of mercury lamp illumination,
respectively (Figure 3d). These results indicate that TPP as
a photoinitiator can improve the photostability of CsPbI3NC–
polymer composite films because TPP can remove oxygen in
polymer gels, reduce chemical damage of NCs during photo-
oxidation, and effectively passivate the surface of NCs.
The TPP-treated CsPbI3NCs have a higher PLQY, good
stability, and lower reabsorption, which is necessary for the
application of LSCs device application in building-integrated
solar windows. A typical LSC consists of a mixture of a plastic
lightguide or a transparent glass plate coated with an emitting
material. The luminophores absorb the sunlight and radiate at
longer wavelengths. It is then guided to the edge by total
internal reflection (TIR) and converted into electricity by
photovoltaic cells mounted on the side of the LSCs. Here,
a layered LSCs device is constructed, coated by a circa 50 mm
layer of the NC–polymer composite onto the high-optical-
quality PMMA glass slab (ca. 2 mm) by using a doctor-blade
deposition technique (Figure 4a,b). It can be clearly seen that
the edge of the slab presents bright emission under weak UV
illumination. The absorption spectrum in Figure 4c of the
CsPbI3-based LSCs device shows a broadband absorption
from the ultraviolet to the near-infrared region.
We further measured the PL spectra of CsPbI3based
LSCs at different optical paths. The integrated PL emission of
the TPP-treated NC–polymer films was maintained at 49 % at
an optical length of 6 cm, while the integrated PL emission of
untreated NC–polymer films decreases to 27% at d=6cm
owing to the larger reabsorption (Figure 4 d). It is well-known
that the optical efficiency (hopt), a key figure of merit (FOM)
for LSCs, is the ratio of the output power (Pout) from the edges
of LSCs devices and the input power (Pin) through the top
surface of the LSCs. When the LSC is coupled with a silicon
PV-cell, the (hopt) can be easily calculated by the following
equation [Eq. (1)]:[12, 14]
Figure 3. a) The molecular structure of several photoinitiators. b) UV/
Vis absorption and PL emission spectra of the different initiator
treated NC–polymer films at illumination time of 0 h. c) The absorb-
ance values at 685 nm of films with TPO, DMPA, HPK, or HMPO as
initiators at different illumination time. d) Time-dependent PLQYs of
different photoinitiators treated NC–polymer films under the strong
mercury lamp without inert gas protection.
Figure 4. a) Preparation of the thin-film LSCs using the doctor-blade
method. b) Photograph of a 5 15 cm CsPbI3NCs-based LSC under
ultraviolet light. c) AM 1.5 G spectrum, UV/Vis absorption, and PL
emission spectra of CsPbI3NCs-based LSCs. d) Integrated PL emission
of the untreated and TPP-treated LSCs at different optical paths.
e) The optical efficiency of the LSCs coupled with silicon cell as
a function of geometric factor (Gfactor). f ) The IVcurves of the Si
solar cell coupled with the edges of an LSC before and after storage
for 2 months in ambient air.
A
ngewandte
Chemi
e
Communications
7740 www.angewandte.org 2020 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2020,59, 7738 –7742
hopt ¼Ilsc=ðIsc GÞð1Þ
where ILSC is the short circuit current from the Si cell coupled
with LSC, Isc is the short circuit current from Si cell under
direct illumination, and Gis the geometric factor, defined as
the ratio of the top area and the edge area.
The LSCs with different sizes were made to test their
optical efficiency. During the course of the experiments, these
LSCs were illuminated by a 1.5 AM global solar simulator
(100 mWcm2). A commercial silicon cell was coupled with
the edge of LSCs. As shown in Figure 4 f, the hopt decreased
exponentially with the increasing of the Gfactors of LSCs
devices, which is also seen in previous reports.[5b,9] The
maximum optical efficiency is 4.9% with a Gfactor of 5.
The optical quantum efficiency hint, which is defined as the
ratio of number of photons emitted from the edge to the
number of photons absorbed by the LSC, is 36% for a 2
2 cm LSCs with G=10. To further evaluate the stability of the
device, we measured the optical efficiency of a 2 2 cm LSCs
before and after 2 months of storage in ambient air (Fig-
ure 4 f). After 2 months of storage, the optical efficiency of
devices only decreased from 3.1% to 2.9 %, still maintaining
93.5% optical efficiency, indicating that these LSCs have
excellent stability.
In summary, we have demonstrated a novel TPP post-
treatment method for efficient and stable thin-film LSCs with
bright and low reabsorption near-infrared emissive perovskite
CsPbI3NC–polymer composite films. These LSCs exhibited
a high absolute PL quantum yield of 99.4 0.4% and long-
term air, water, and light stability. These thin-film LSCs
exhibit an impressive optical conversion efficiency of 3.1%
and optical quantum efficiency of 36 % at a geometric factor
of 10. This work may lay a foundation for efficient and stable
LSCs based on low-cost perovskite NCs.
Acknowledgements
This work was supported by Natural Science Foundation of
Jiangsu Province (Grant No. BZ2018008) and Shuangchuang
Program of Jiangsu Province.
Conflict of interest
The authors declare no conflict of interest.
Keywords: lead halides · luminescent solar concentrators ·
nanocrystals · perovskites · polymer composites
How to cite: Angew. Chem. Int. Ed. 2020,59, 77387742
Angew. Chem. 2020,132, 7812– 7816
[1] a) J. H. Yu, S.-H. Kwon, Z. Petrs
ˇek, O. K. Park, S. W. Jun, K.
Shin, M. Choi, Y. I. Park, K. Park, H. B. Na, N. Lee, D. W. Lee,
J. H. Kim, P. Schwille, T. Hyeon, Nat. Mater. 2013,12, 359; b) F.
Wang, R. Deng, J. Wang, Q. Wang, Y. Han, H. Zhu, X. Chen, X.
Liu, Nat. Mater. 2011,10, 968.
[2] a) R. Wang, X. Wu, K. Xu, W. Zhou, Y. Shang, H. Tang, H. Chen,
Z. Ning, Adv. Mater. 2018,30, 1704882; b) Y. Cho, P. Giraud, B.
Hou, Y. W. Lee, J. Hong, S. Lee, S. Pak, J. Lee, J. E. Jang, S. M.
Morris, Adv. Energy Mater. 2018,8, 1870012; c) Y. Zhang, G. Wu,
C. Ding, F. Liu, Y. Yao, Y. Zhou,C. Wu, N. Nakazawa, Q. Huang,
T. Toyoda, J. Phys. Chem. Lett. 2018,9, 3598.
[3] a) H. Shen, Q. Lin, H. Wang, L. Qian, Y. Yang, A. Titov, J.
Hyvonen, Y. Zheng, L. S. Li, ACS Appl. Mater. Interfaces 2013,5,
12011; b) Z. Shi, L. Ying, Y. Zhang, Y. Chen, X. Li, W. Di, T. Xu,
C. Shan, G. Du, Nano Lett. 2017,17, 313; c) K. Yang, F. Li, Y.
Liu, Z. Xu, Q. Li, K. Sun, L. Qiu, Q. Zeng, Z. Chen, W. Chen,
ACS Appl. Mater. Interfaces 2018,10, 27374.
[4] a) S. A. Mcdonald, G. Konstantatos, S. Zhang, P. W. Cyr, E. J.
Klem, L. Levina, E. H. Sargent, Nat. Mater. 2005,4, 138; b) L. Li,
L. Gu, Z. Lou, Z. Fan, G. Shen, ACS Nano 2017,11, 4067; c) T.
Shen, B. Li, K. Zheng, T. Pullerits, G. Cao, J. Tian, J. Phys. Chem.
Lett. 2018,9, 3285.
[5] a) Y. Zhou, H. Zhao, D. Ma, F. Rosei, Chem. Soc. Rev. 2018,47,
5866; b) Y. Zhou, D. Benetti, Z. Fan, H. Zhao, D. Ma, A. O.
Govorov, A. Vomiero, F. Rosei, Adv. Energy Mater. 2016,6,
1501913; c) H. Zhao, D. Benetti, L. Jin, Y. Zhou, F. Rosei, A.
Vomiero, Small 2016,12, 5354; d) H. Li, K. Wu, J. Lim, H.-J.
Song, V. I. Klimov, Nat. Energy 2016,1, 16157; e) C. Li, W. Chen,
D. Wu, D. Quan, Z. Zhou, J. Hao, J. Qin, Y. Li, Z. He, K. Wang,
Sci. Rep. 2015,5, 17777.
[6] J. Y. Kim, O. Voznyy, D. Zhitomirsky, E. H. Sargent, Adv. Mater.
2013,25, 4986.
[7] Y. Zhou, D. Benetti, T. Xin, J. Lei, Z. M. Wang, D. Ma, H. Zhao,
F. Rosei, Nano Energy 2018,44, 378.
[8] F. Meinardi, S. Ehrenberg, L. Dhamo, F. Carulli, M. Mauri, F.
Bruni, R. Simonutti, U. Kortshagen, S. Brovelli, Nat. Photonics
2017,11, 177.
[9] C. S. Erickson, L. R. Bradshaw, S. Mcdowall, J. D. Gilbertson,
D. R. Gamelin, D. L. Patrick, ACS Nano 2014,8, 3461.
[10] H. Xiangmin, K. Ruidan, Z. Yongyou, D. Luogen, Z. Haizheng,
Z. Bingsuo, S. Li-Jie, Opt. Express 2015,23, 858.
[11] F. Meinardi, H. McDaniel, F. Carulli, A. Colombo, K. A.
Velizhanin, N. S. Makarov, R. Simonutti, V. I. Klimov, S.
Brovelli, Nat. Nanotechnol. 2015,10, 878.
[12] I. Coropceanu, M. G. Bawendi, Nano Lett. 2014,14, 4097.
[13] a) A. Dutta, S. K. Dutta, S. Das Adhikari, N. Pradhan, Angew.
Chem. Int. Ed. 2018,57, 9083; Angew. Chem. 2018,130, 9221;
b) M. V. Kovalenko, L. Protesescu, M. I. Bodnarchuk, Science
2017,358, 745.
[14] K. Nikolaidou, S. Sarang, C. Hoffman, B. Mendewala, H.
Ishihara, J. Q. Lu, B. Ilan, V. Tung, S. Ghosh, Adv. Opt. Mater.
2016,4, 2126.
[15] a) X. Luo, T. Ding, X. Liu, Y. Liu, K. Wu, Nano Lett. 2019,19,
338; b) H. Zhao, Y. Zhou, D. Benetti, D. Ma, F. Rosei, Nano
Energy 2017,37, 214.
[16] F. Meinardi, Q. A. Akkerman, F. Bruni, S. Park, M. Mauri, Z.
Dang, L. Manna, S. Brovelli, ACS Energy Lett. 2017,2, 2368.
[17] M. Wei, F. P. G. de Arquer, G. Walters, Z. Yang, L. N. Quan, Y.
Kim, R. Sabatini, R. Quintero-Bermudez, L. Gao, J. Z. Fan, F.
Fan, A. Gold-Parker, M. F. Toney, E. H. Sargent, Nat. Energy
2019,4, 197.
[18] J. Tong, J. Luo, L. Shi, J. Wu, L. Xu, J. Song, P. Wang, H. Li, Z.
Deng, J. Mater. Chem. A 2019,7, 4872.
[19] a) F. Liu, Y. Zhang, C. Ding, S. Kobayashi, T. Izuishi, N.
Nakazawa, T. Toyoda, T. Ohta, S. Hayase, T. Minemoto, K.
Yoshino, S. Dai, Q. Shen, ACS Nano 2017,11, 10373; b) C. Lu, H.
Li, K. Kolodziejski, C. Dun, W. Huang, D. Carroll, S. M. Geyer,
Nano Res. 2018,11, 762; c) H. Wang, N. Sui, X. Bai, Y. Zhang, Q.
Rice, F. J. Seo, Q. Zhang, V. L. Colvin, W. W. Yu, J. Phys. Chem.
Lett. 2018,9, 4166.
[20] a) M. Lu, X. Zhang, Y. Zhang, J. Guo, X. Shen, W. W. Yu, A. L.
Rogach, Adv. Mater. 2018,30, 1804691; b) J. Pan, Y. Shang, J.
Yin, M. De Bastiani, W. Peng, I. Dursun, L. Sinatra, A. M. El-
A
ngewandte
Chemi
e
Communications
7741Angew. Chem. Int. Ed. 2020,59, 7738 –7742 2020 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.angewandte.org
Zohry, M. N. Hedhili, A. H. Emwas, O. F. Mohammed, Z. Ning,
O. M. Bakr, J. Am. Chem. Soc. 2018,140, 562.
[21] C. Bi, S. V. Kershaw, A. L. Rogach, J. Tian, Adv. Funct. Mater.
2019,29, 1902446.
[22] T. J. Mao, R. J. Eldred, J. Polym. Sci. Part A 1967,5, 1741.
Manuscript received: September 11, 2019
Revised manuscript received: November 18, 2019
Accepted manuscript online: January 30, 2020
Version of record online: March 6, 2020
A
ngewandte
Chemi
e
Communications
7742 www.angewandte.org 2020 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2020,59, 7738 –7742
... Embedding PNCs into polymer is an effective strategy to enhance the PNCs stability and polymer can endow the PNCs with other positive effects based on different structure and functional groups, such as surface defect passivation 10 , facilitating the charge separation and transport 11 , assisted self-assembly and controlled morphology [12][13][14] , and excellent processability, stretchability, and mechanical property in the form of nanocomposites 15 . Therefore, PNCs/polymer nanocomposite is promising for displays [15][16][17] , luminescent solar concentrators 18,19 , scintillator 20,21 , lighting [22][23][24] and hybrid photovoltaic device 25,26 . The uniform distribution of PNCs in polymer matrix is critical to the properties of the nanocomposites and the aggregation of PNCs induced by high surface energy has a severe influence on the performance of related applications [27][28][29][30][31] . ...
Article
Full-text available
Perovskite nanocrystals (PNCs)/polymer nanocomposites can combine the advantages of each other, but extremely few works can achieve the fabrication of PNCs/polymer nanocomposites by bulk polymerization. We originally adopt a two-type ligand strategy to fabricate bulk PNCs/polystyrene (PS) nanocomposites, including a new type of synthetic polymerizable ligand. The CsPbCl 3 PNCs/PS nanocomposites show extremely high transparency even the doping content up to 5 wt%. The high transparency can be ascribed to the Rayleigh scattering as the PNCs distribute uniformly without obvious aggregation. Based on this behavior, we first exploit the potential of PNCs to serve as scatters inside light guided plate (LGP), whose surface illuminance and uniformity can be improved, and this new kind of LGP is compatible with the advanced liquid crystal display technology. Thanks to the facile composition adjustment of CsPbCl x Br 3- x (1 ≤ x ≤ 3) PNCs, the Rayleigh scattering behavior can also be adjusted so as to the performance of LGP. The best-performing 5.0-inch LGP based on CsPbCl 2.5 Br 0.5 PNCs/PS nanocomposites shows 20.5 times higher illuminance and 1.8 times higher uniformity in display than the control. The LGP based on PNCs/PS nanocomposite exhibits an enormous potential in commercialization no matter based on itself or combined with the LGP-related technology.
... [4] In order to explore earth-abundant and facilely assembled SSL materials, researchers have paid more attention to design new luminescent materials including all-inorganic semiconductor nanocrystals, metal-organic complexes, organic/ polymeric emitters, and organic-inorganic hybrid metal halide perovskites. [5][6][7][8][9] Recently, the newly emerged 0D organic-inorganic hybrid halide perovskites have been widely applied as most promising SSL materials in future lighting and display field owing to multiple advantages of diversified structural architectures, excellent luminescent performance including tunable emitting color and high photoluminescence quantum yield (PLQY), etc. [10][11][12][13][14][15][16] Significantly, versatile organic cations and numerous alternative metal ions (In 3 + , Sb 3 + , Mn 2 + , Cu + , Zn 2 + , etc) endow 0D hybrid halides with diversified structural architectures and fascinating photoluminescence (PL) performance, which provide more opportunities to modulate the PL properties. [17][18][19] Especially, 0D hybrid Sb 3 + -based halides represent one of the most favorable luminescent materials with abundant coordination configurations, higher PLQYs and oxidation resistance abilities comparing with other counterparts. ...
Article
Full-text available
Recently, zero‐dimensional (0D) hybrid metal halides have attracted intensive attention with wide applications in solid‐state lighting and display diodes. Herein, by using a facile wet‐chemistry method, we prepared one new 0D hybrid antimony halide of [HMHQ]2SbCl5 ⋅ 2H2O (MHQ=2‐methyl‐8‐hydroxyquinoline) based on the discrete [SbCl5]²⁻ unit. Remarkably, the bulk crystals of [HMHQ]2SbCl5 ⋅ 2H2O exhibit strong cyan light emission with a promising photoluminescence quantum yield (PLQY) of 18.92%. Systematical studies disclose that the cyan emission is mainly derived from the radiative recombination within conjugated organic cation. Benefiting from the promising luminescent performance, this 0D antimony halide can be utilized as an excellent down‐conversion light emitting luminescent material to assemble white light‐emitting diodes with high color rendering index (CRI) of 90.2.
... 1 Among their most notable characteristics include (i) their relatively easy preparation, (ii) narrow width at half-maximum photoluminescence (PL) peak, 2 (iii) an adjustable/modifiable surface chemistry, 3,4 (iv) a notable PL quantum yield (PLQY) of up to 100%, 5−8 and (v) a modulable band gap (E g ) by varying the PNC size due to the quantum confinement effect, composition of A-site cations (e.g., formamidinium cation, Cs + ), or X-site halides in the ABX 3 -type perovskites, 9−11 taking into account that phase photosegregation in mixed-halide PNCs is restricted in comparison with their bulk counterparts. 12 These intrinsic features have been exploited to fabricate efficient PNC solar cells 13 with a high photoconversion efficiency of up to 18.1%, 14 light-emitting diodes (LEDs) with external quantum efficiencies (EQEs) surpassing 20%, 15,16 scintillators, 17 photodetectors, 18 light amplifiers, 19 lasers, 20 and solar concentrators 21 among others. Despite the benign defect physics of halide perovskites, the PLQY is limited by the defective structure, which in addition makes them prone to degradation. ...
Article
Full-text available
Current efforts to reduce the density of structural defects such as surface passivation, doping, and modified synthetic protocols have allowed us to grow high-quality perovskite nanocrystals (PNCs). However, the role of the purity of the precursors involved during the PNC synthesis to hinder the emergence of defects has not been widely explored. In this work, we analyzed the use of different crystallization processes of PbX2 (X = Cl– or I–) to purify the chemicals and produce highly luminescent and stable CsPbCl3–xBrx and CsPbI3 PNCs. The use of a hydrothermal (Hyd) process to improve the quality of the as-prepared PbCl2 provides blue-emitting PNCs with efficient ligand surface passivation, a maximum photoluminescence quantum yield (PLQY) of ∼ 88%, and improved photocatalytic activity to oxidize benzyl alcohol, yielding 40%. Then, the hot recrystallization of PbI2 prior to Hyd treatment led to the formation of red-emissive PNCs with a PLQY of up to 100%, long-term stability around 4 months under ambient air, and a relative humidity of 50–60%. Thus, CsPbI3 light-emitting diodes were fabricated to provide a maximum external quantum efficiency of up to 13.6%. We claim that the improvement of the PbX2 crystallinity offers a suitable stoichiometry in the PNC structure, reducing nonradiative carrier traps and so maximizing the radiative recombination dynamics. This contribution gives an insight into how the manipulation of the PbX2 precursor is a profitable and potential alternative to synthesize PNCs with improved photophysical features by making use of defect engineering.
... In the past fifty years, scientists have demonstrated the LSC device using different types of organic dye molecules (Rafiee et al., 2019), upconverted materials (Ha et al., 2018;Nam et al., 2020), rare earth complexes (Day et al., 2019;Wang et al., 2011), perovskite nanocrystals Wu et al., 2020), fluorescent proteins (Sadeghi et al., 2019) as the luminescent species (Mazzaro and Vomiero, 2018;Meinardi et al., 2017a) including green alternatives (Z. . Actual number of reports are huge compared to the references cited here. ...
Article
A novel approach to creating a multifunctional smart window that combines electricity generation with transparency modulation is developed. The proposed smart window is based on a mechanoresponsive luminescent solar concentrator (LSC) composed of organic fluorophores (coumarin 6) impregnated into a compressible polymer matrix (polydimethylsiloxane, PDMS) that is surface‐modified on one side. In its initial state, the LSC is optically transparent but becomes increasingly opaque upon the application of compressive strain to the PDMS matrix, resulting in a higher optical shielding property and eventually becoming translucent. The smart window's degree of specular transmission can be reversibly adjusted between 20% and 84%, making it suitable for use as a switchable privacy window. Additionally, the LSC‐photovoltaic (PV) window maintains a high power conversion efficiency (PCE) even when in a translucent state with the privacy function enabled. Under AM1.5G solar irradiation (1 sun, 100 mW cm ² ), the LSC‐PV module achieves PCEs of 1.05–1.48% and optical efficiencies ( η opt ) of 8.6–12.0%, depending on the strain. This technology holds significant promise for building‐integrated photovoltaics and smart window applications, such as privacy‐protecting windows, daylight‐controlling windows, power‐generating antinoise barriers, and transparent solar cells.
Article
Herein, we report two hybrid zero‐dimensional (0D) lead‐free metal oxide halides of [PPip]Sb 2 Cl 6 O ( 1 ) ([PPip] = 1‐(2‐Pyridyl)piperazine) and [PmPip]Sb 2 Cl 6 O ( 2 ) ([PmPip] = 1‐(2‐Pyrimidyl)piperazine). Compounds 1 and 2 are able to exhibit narrow blue light emissions excited by UV light with photo‐luminescence quantum yields (PLQYs) of 25.49% and 24.71%, respectively. Through systematical characterizations, these blue light emissions can be attributed to the radiative transition in conjugated organic cations. Benefiting from the highly efficient intrinsic blue emissions, compound 2 can be used as blue fluorophor to fabricate liquid crystal display. This work provides a new to assembly blue emitter through by virtue of the optical active organic cations in hybrid materials.
Article
Full-text available
A surface engineering strategy aimed at improving the stability of CsPbI3 perovskite quantum dots (QDs) both in solution and as films is demonstrated, by performing partial ligand exchange with a short chain ligand, 2‐aminoethanethiol (AET), in place of the original long chain ligands, oleic acid (OA) and oleylamine (OAm), used in synthesis. This results in the formation of a compact ligand barrier around the particles, which prevents penetration of water molecules and thus degradation of the films and, in addition, at the same time improves carrier mobility. Moreover, the AET ligand can passivate surface traps of the QDs, leading to an enhanced photoluminescence (PL) efficiency. As a result, AET‐CsPbI3 QDs maintain their optical performance both in solution and as films, retaining more than 95% of the initial PL intensity in water after 1 h, and under ultraviolet irradiation for 2 h. Photodetectors based on the AET‐CsPbI3 QD films exhibit remarkable performance, such as high photoresponsivity (105 mA W⁻¹) and detectivity (5 × 10¹³ Jones at 450 nm and 3 × 10¹³ Jones at 700 nm) without an external bias. The photodetectors also show excellent stability, retaining more than 95% of the initial responsivity in ambient air for 40 h without any encapsulation.
Article
Full-text available
In luminescent solar concentrator (LSC) systems, broadband solar energy is absorbed, down-converted and waveguided to the panel edges where peripheral photovoltaic cells convert the concentrated light to electricity. Achieving a low-loss LSC requires reducing the reabsorption of emitted light within the absorbing medium while maintaining high photoluminescence quantum yield (PLQY). Here we employ layered hybrid metal halide perovskites—ensembles of two-dimensional perovskite domains—to fabricate low-loss large-area LSCs that fulfil this requirement. We devised a facile synthetic route to obtain layered perovskite nanoplatelets (PNPLs) that possess a tunable number of layers within each platelet. Efficient ultrafast non-radiative exciton routing within each PNPL (0.1 ps⁻¹) produces a large Stokes shift and a high PLQY simultaneously. Using this approach, we achieve an optical quantum efficiency of 26% and an internal concentration factor of 3.3 for LSCs with an area of 10 × 10 cm², which represents a fourfold enhancement over the best previously reported perovskite LSCs. © 2019, The Author(s), under exclusive licence to Springer Nature Limited.
Article
The high-performance light-management slabs made of low cost, highly emissive, and ultra-stable nanocrystals (NCs)-polymer composites are desirable for the application of large-area luminescent solar concentrators (LSCs). However, although the reported photoluminescence (PL) quantum yield (QY) of NCs in solution is up to 90%, the PL QY of the NCs in the polymer matrix is usually low, which limits the performance of the LSCs. Herein, we demonstrate a new strategy for the synthesis of formamidinium lead bromide (FAPbBr3) NCs via a room-temperature solvent-induced reprecipitation using dicarboxylic acids as ligands and the preparation of NC-polymer composite slabs for LSCs. Due to the strong binding of dicarboxylic acids, the as-synthesized decanedioic acid (DA)-capped FAPbBr3 NCs displayed high PL QY (90±5%), increased chemical yield, and improved stability, as compared to the monocarboxylic oleic acid (OA)-capped NCs (PL QY of 80±5%) synthesized in the same conditions. Furthermore, in the presence of toluene as the solvent, these DA-capped NCs exhibited good compatibility with polystyrene (PS), so the NCs-PS slurry has an appropriate viscosity. Therefore, it is convenient to deposit the slurry onto commercially available polymethyl methacrylate (PMMA) slabs with a standard doctor-blade. The optimized NCs-PS-PMMA slabs with PL QY of 92±5% were sustained for over 1000 hours under high temperature (60oC) and high humidity (relative humidity, RH=90%) environments. The results show that the perovskite NCs-based slabs developed in this study has the advantages of low cost, highly emissivity, and good stability, and may contribute to the field of LSCs.
Article
A method is proposed to improve the photo/electroluminescence efficiency and stability of CsPbI3 perovskite nanocrystals (NCs) by using SrCl2 as a co‐precursor. The SrCl2 is chosen as the dopant to synthesize the CsPbI3 NCs. Because the ion radius of Sr2+ (1.18 Å) is slightly smaller than that of Pb2+ (1.19 Å) ions, divalent Sr2+ cations can partly replace the Pb2+ ions in the lattice structure of perovskite NCs and cause a slight lattice contraction. At the same time, Cl− anions from SrCl2 are able to efficiently passivate surface defect states of CsPbI3 nanocrystals, thus converting nonradiative trap states to radiative states. The simultaneous Sr2+ ion doping and surface Cl− ion passivation result in the enhanced photoluminescence quantum yield (up to 84%), elongated emission lifetime, and improved stability. Sr2+‐doped CsPbI3 NCs are employed to produce light‐emitting devices with a high external quantum yield of 13.5%. SrCl2 is introduced as a co‐precursor in the synthesis of CsPbI3 perovskite nanocrystals to realize their simultaneous Sr2+ cation doping and surface Cl− anion passivation. The stability of nanocrystals is improved, and light‐emitting devices with a high external quantum efficiency of 13.5% are realized.
Article
Inorganic lead halide perovskite quantum dots (PQDs), especially red emission PQDs, are well known to easily lose their luminescence emission with time, which shows from strong emission of fresh PQDs to no emission of aged PQDs. Here, we demonstrate that trioctylphosphine (TOP) can effectively and instantly recover the luminescence emission of aged red PQDs, making the “dead” PQDs “reborn”. Furthermore, TOP also works to improve the emission intensity of freshly synthesized PQDs. In this process, TOP does not make any detectable structural changes to PQDs. Besides, TOP can effectively enhance the stability of PQDs against long term storage, temperature, UV irradiation, and polar solvents. This unusual emission recovery and stability enhancement by TOP shall promote the understanding of particle surface condition and the development of PQD devices.
Article
Luminescent solar concentrators (LSCs) can serve as large-area sunlight collectors, are suitable for applications in high-efficiency and cost-effective photovoltaics (PVs), and provide adaptability to the needs of architects for building-integrated PVs, which makes them an attractive option for transforming buildings into transparent or non-transparent electricity generators. Compared with traditional organic dyes, colloidal semiconducting quantum dots (QDs) are excellent candidates as emitters for LSCs because they exhibit wide size/shape/composition-tunable absorption spectra ranging from ultraviolet to near infrared, significantly overlapping with the solar spectrum. They also feature narrow emission spectra, high photoluminescence quantum yields, high absorption coefficients, solution processability and good photostability. Most importantly, QDs can be engineered to provide a minimal overlap between absorption and emission spectra, which is key to the realization of large-area LSCs with largely suppressed reabsorption energy losses. In this review article, we will first present and discuss the working principle of LSCs, the synthesis of colloidal QDs using wet-chemistry approaches, the optical properties of QDs, their band alignment and the intrinsic relationship between the band energy structure and optical properties of QDs. We focus on emerging architectures, such as core/shell QDs. We then highlight recent progress in QD-based LSCs and their anticipated applications. We conclude this review article with the major challenges and perspectives of LSCs in future commercial technologies.
Article
Lead selenide (PbSe) colloidal quantum dots (CQDs) are considered to be a strong candidate for high-efficiency colloidal quantum dot solar cells (CQDSCs) due to its efficient multiple exciton generation. However, currently, even the best PbSe CQDSCs can only display open-circuit voltage (Voc) about 0.530 V. Here, we introduce a solution-phase ligand exchange method to prepare PbI2-capped PbSe (PbSe-PbI2) CQD inks, and for the first time, the absorber layer of PbSe CQDSCs was deposited in one step by using this PbSe-PbI2 CQD inks. One-step-deposited PbSe CQDs absorber layer exhibits fast charge transfer rate, reduced energy funneling and low trap assisted recombination. The champion large-area (active area is 0.35 cm2) PbSe CQDSCs fabricated with one-step PbSe CQDs achieve a power conversion efficiency (PCE) of 6.0%, and a Voc of 0.616 V, which is the highest Voc among PbSe CQDSCs reported to date.
Article
Quantum dot (QDs)-based photodetectors are capable of a broad spectrum from ultraviolet to near infrared (UV-NIR), have low dark current, and can operate at high temperatures, however, so far the detectivity of QDs-based photodetector has been low. Herein, the ternary alloyed CdSexTe1-x colloidal QDs trap-passivated by iodide-based ligands (TBAI) are developed as building blocks for UV-NIR photodetector with a configuration of n-TiO2/QDs/p-Spiro-OMeTAD. Both the less surface traps and much high loading of QDs are obtained by in situ ligand exchange with TBAI. The device is sensitive to a broad wavelength range covering the entire UV-NIR region (300~850 nm), showing an excellent photo-responsivity of 53 mA/W, a fast response time of << 0.02s and remarkably high-detectivity values of 8×1013 Jones at 450 nm and 1×1013 at 800 nm without external bias voltage, respectively. Such performance is superior to what has been reported earlier for QD-based photodetectors. The photodetector exhibits excellent stability keeping 98% of photoelectric responsivity after 2-month illumination in air even without encapsulation. In addition, the semitransparent device is successfully fabricated using Ag nanowires/polyimide transparent substrate. Such self-powered photodetectors with fast response speed, stable, broadband response are expected to function under broad range of environmental conditions.