In Situ Tuning the Reactivity of Selenium Precursor To Synthesize

Jan 7, 2018 - School of Optical and Electronic Information, Huazhong University of Science and Technology, 1037 Luoyu Road, Wuhan, Hubei 430074, ...
33 downloads 3 Views 5MB Size
Article Cite This: Chem. Mater. 2018, 30, 982−989

pubs.acs.org/cm

In Situ Tuning the Reactivity of Selenium Precursor To Synthesize Wide Range Size, Ultralarge-Scale, and Ultrastable PbSe Quantum Dots Linyuan Lian,† Yong Xia,† Changwang Zhang,† Bing Xu,‡ Lei Yang,‡ Huan Liu,† Daoli Zhang,† Kai Wang,‡ Jianbo Gao,§ and Jianbing Zhang*,†,⊥ †

School of Optical and Electronic Information, Huazhong University of Science and Technology, 1037 Luoyu Road, Wuhan, Hubei 430074, China ‡ Department of Electrical and Electronic Engineering, Southern University of Science and Technology, Shenzhen, Guangdong 518055, China § Department of Physics and Astronomy, University of Clemson, Clemson, South Carolina 29634, United States ⊥ Shenzhen R & D Center of Huazhong University of Science and Technology, Shenzhen, Guangdong 518057, China S Supporting Information *

ABSTRACT: PbSe quantum dots (QDs) have shown outstanding optoelectronic properties due to their extremely strong quantum confinement. However, the applications of PbSe QDs are substantially limited by their instability in air. Here, we developed a simple synthesis for PbSe QDs via in situ tuning the reactivity of selenium precursor. Because of in situ chloride passivation, the as prepared PbSe QDs showed excellent stability in air, demonstrated by the retention of absorption features and photoluminescence quantum yields (PL QYs) for different sizes after heated at 80 °C in air. Furthermore, the ligand exchanged electronic coupled PbSe QD thin films also had excellent stability in air even for large particle sizes. In addition, the PbSe QDs showed high PL QYs due to hybrid organic (oleate) and inorganic (Cl−) passivation. Monodispersive large (1st exciton peak > 2100 nm) and extremely small (1st exciton peak < 750 nm) PbSe QDs were achieved by careful control of particle growth. As a heating up method, this new synthesis was easily scaled up to produce 23.5 g PbSe QDs from one batch of reaction. The versatility of this new synthetic strategy was demonstrated by the synthesis of other metal selenide QDs such as CdSe and ZnSe QDs.



INTRODUCTION

PbS QD solar cells. (3) There are fewer intragap states in electronically coupled PbSe than PbS QD films.11 Additionally, favorable band-like transport has been observed in PbSe QD films but not yet in PbS QD films.12 (4) PbSe QDs can form epitaxial connections between well-oriented proximal nanoparticles (2D superlattices),13,14 while this kind of assembling is not achieved in PbS QDs. The confined-but-connected PbSe films show much better carrier mobility15 and higher optical absorption cross sections.16 (5) PbSe QDs have lower bulk bandgap and stronger quantum confinement than PbS QDs, and as a result, the bandgap of PbSe QDs can be tuned in a larger range, which will find a wider range of applications. Therefore, PbSe QDs should exhibit better device performance than PbS QDs and would become ideal QD materials if the air-sensitivity can be addressed. However, large scale, wide range size synthesis while maintaining air-stability are the main challenges for PbSe QD

Lead chalcogenide (PbS, PbSe, and PbTe) quantum dots (QDs) have become important optoelectronic and electronic materials due to strong quantum confinement, tunable bandgap in a large range, ease of synthesis, solution processability, and efficient multiple exciton generation (MEG).1−3 Among the lead chalcogenide QDs, PbS are the most studied materials due to their acceptable stability and have been applied in various devices.1,4,5 Compared to PbS, PbSe QDs are less studied due to their air-sensitivity. However, PbSe QDs have excellent intrinsic optoelectronic properties, which might lead to better device performance. (1) The exciton Bohr radius of PbSe is 46 nm, which is more than twice that of PbS (20 nm).6 As a result, PbSe QDs exhibit much stronger quantum confinement than PbS QDs, which should induce more leakage of wave function in PbSe QDs and in turn stronger electronic coupling between QDs and better charge carrier transport in PbSe QD films.7 (2) PbSe QDs have better MEG performance than PbS QDs. First, ultrafast spectroscopic experiments show that PbSe QDs have higher MEG efficiency.8 Second, QD solar cells with EQE above 100% have been observed in PbSe9,10 QD devices but not yet in © 2018 American Chemical Society

Received: November 16, 2017 Revised: January 3, 2018 Published: January 7, 2018 982

DOI: 10.1021/acs.chemmater.7b04825 Chem. Mater. 2018, 30, 982−989

Article

Chemistry of Materials

(OLA, tech. grade, 70%), diphenylphosphine (DPP, 95%), and selenium powder (99.999%) were purchased from Aladdin. Tetrachloroethylene (TCE, ≥98.5%), hexane (≥97%), and acetone (≥99.5%) were purchased from Sinopharm Chemical Reagent. All the chemicals were used as received. TOPSe and DPPSe (selenium powder dissolved in TOP and DPP, respectively) were used as the molecular precursors. DPPSe was made by heating the DPP and Se powder at 120 °C for 30 min. Synthesis of PbSe QDs. In a typical synthesis, OLA (36 mL) and 9 mmol of PbCl2 were degassed under vacuum at 80 °C and heated to 140 °C under nitrogen, and the temperature was maintained for 30 min until a white and turbid solution achieved. Then the suspension was allowed to cool to 30 °C, 110 μL of 1 M DPPSe diluted in 2 mL of ODE, 0.57 mL of 1 M TOPSe and 110 μL of DPP were injected sequentially to the lead precursor solution at 30 °C. Next, the reaction solution was reheated to 160 °C under vigorous stirring for the growth of the QDs. At various temperatures, 30 μL reaction solutions were withdrawn for the absorption measurements. The growth solution was maintained at 160 °C for 15 min, and aliquots were withdrawn at different time intervals. The aliquots were dispersed in hexane with OA, washed one time using acetone, and dispersed in 3 mL of TCE for the measurement of absorption spectra. Second injection of molecular precursors was needed to obtain larger PbSe QDs. After the growth solution was maintained at 160 °C for 15 min, 0.57 mL of 1 M TOPSe and 110 μL of DPP diluted in 2 mL of ODE were injected sequentially to the growth solution, and the solution was maintained at 160 °C for another 15 min, which resulted in PbSe QDs with the first exciton peak beyond 2000 nm. If the volume of the second injection reagents doubled, 1.14 mL of 1 M TOPSe and 220 μL of DPP, PbSe QDs with the first exciton peak beyond 2100 nm were obtained. Purification of PbSe QDs. The reaction was quenched by a water bath when the desired size was achieved, and 20 mL of hexane and 20 mL of OA were added at 70 and 40 °C, respectively, followed by vigorous stirring for 10 min. The purpose of addition of OA is replacing the weakly bound OLA. The unreacted lead precursors were precipitated from the raw solution by centrifugation after the addition of hexane and OA, and the precipitate was discarded. The supernatant solution was collected and precipitated by adding acetone. After centrifugation, the supernatant was discarded and the precipitated PbSe QDs was collected. Synthesis of CdSe and ZnSe QDs. The synthesis of CdSe and ZnSe QDs followed the same manner as that of PbSe QDs. In a typical synthesis of CdSe QDs, 0.256 g of CdO, 1.57 g of OA, and 12.5 g of ODE were loaded in a 50 mL round-bottom three-neck flask, and then the mixture was degassed under vacuum at 80 °C and heated to 260 °C under nitrogen and the temperature was maintained for 20 min until the solution became colorless. Then the solution was allowed to cool to 30 °C, 110 μL of 1 M DPPSe diluted in 2 mL of ODE, 0.57 mL of 1 M TOPSe and 110 μL of DPP were injected sequentially to the cadmium precursor solution at 30 °C. Next, the reaction solution was reheated to 270 °C under vigorous stirring for the growth of the QDs. At various temperatures, reaction solutions were withdrawn for the absorption measurements. The synthesis of ZnSe QDs is the same with that of CdSe QDs except the Cd precursor (CdO and OA) was replaced by zinc stearate and the temperature was raised to 290 °C. Characterization. Optical absorption spectra were collected using a Shimadzu UV-3600 plus spectrophotometer. PL QYs were measured by an Ocean Optics QE65 Pro spectrometer equipmented with a Labsphere integrating sphere of 3.3-in. innerdiameter. TEM images were obtained using a FEI Technai G2 20 microscope with a LaB6 filament operated at 200 kV. XPS data were obtained on a Kratos AXISULTRA DLD and the sample was prepared by drop casting QD solution on ITO coated glass. X-ray diffraction (XRD) patterns were recorded on a XRD-7000S diffractometer (Shimadzu). Calculation of Dot Size and Particle Concentration. The diameter and concentration of QDs were calculated using the relationship between the optical bandgap and QD size and the relationship between QD diameter and molar extinction coefficient at the first exciton peak given by Yu et al.26

device applications. The PbSe QDs degrade gradually demonstrated by the blueshift of absorption and photoluminescence (PL) peaks and quenching of PL intensity.17,18 X-ray photoelectron spectroscopy (XPS) study shows that up to 50% of particle volume is transformed into PbO, SeO2, or PbSeO3 within 24 h when PbSe QDs solution is exposed to air under ambient conditions.18 Therefore, PbSe QDs would become ideal QD materials if above challenges can be addressed. To improve the stability of PbSe QDs, different postsynthetic strategies were developed. Pietryga et al. coated PbSe QDs with CdSe forming PbSe/CdSe QDs.19 Although the core/shell structure improves the stability substantially, the CdSe shell with a wider bandgap is a barrier for carrier transport. Bae et al. treated PbSe QDs using molecular chlorine (Cl2) to replace the surface Se with Cl atoms.20 The chloride passivated PbSe QDs dispersed in solvents show excellent stability indicating halide passivation is an effective way to stabilize PbSe QDs. Jeong et al. improved the postsynthetic halide passivation by injecting halide salts to PbSe QDs growth solution after the formation of nanoparticles.21 Additionally, in situ passivation was also derived by including capping ligands in the nucleation and growth of PbSe QDs. Jeong et al. introduced phosphonic acid in the synthesis of PbSe QDs.22 The P−O− moieties on the surface render the PbSe QDs stable in solvents and in electronically coupled films. However, the long chain phosphonic acids remained on the QD surface after ligand exchange would limit the carrier transport in the PbSe QD films. Recently, Beard and Zhang reported cation exchange synthesis for stable halide passivated PbSe QDs;23,24 however, combination of cation exchange and quantized Ostwald ripening is needed for monodispersive large sized PbSe QDs.25 Therefore, a simple and large scale synthesis for monodispersive PbSe QDs with tunable sizes in a large range is highly desirable. In the present work, we developed a new method for the synthesis of ultrastable, wide range dot size, scale-up to >20 g PbSe QDs. The complex of PbCl2 and oleylamine (OLA) was adopted as the Pb precursor for in situ chloride passivation. Diphenylphosphine selenide (DPPSe) and trioctylphosphine selenide (TOPSe) were used as the Se precursors. The Pb and Se precursors were mixed at room temperature and the growth of PbSe QDs was driven by heating the growth solution. The TOPSe with low reactivity was gradually converted to reactive DPPSe as the increase of temperature, promoting the growth and maintaining a narrow size distribution. Because of the in situ chloride passivation, the as prepared PbSe QDs showed high stability in air in the form of powder and high photoluminescence quantum yields (PL QYs). Furthermore, stable and electronically coupled PbSe QD thin films with the firstexciton peak above 2000 nm were achieved for the first time. The size can be controlled in a large range, including extremely small dots and large dots with firstexciton peak above 2100 nm, via temperature, the ratio of TOPSe:DPPSe, or injection of additional Se precursor. Slow Ostwald ripening was observed in the new synthesis, facilitating the accurate size control of the PbSe QDs. As a heating up method, the new synthesis was easily scaled up to produce 23.5 g monodispersive PbSe QDs from one batch of reaction. The versatility of this new strategy was demonstrated by the applications in the synthesis of CdSe and ZnSe QDs.



EXPERIMENTAL SECTION

Chemicals. PbCl2 (99.999%), CdO (≥99.95%), zinc stearate (ZnO 12.5−14%), oleic acid (OA, tech. grade, 90%), 1-octadecene (ODE, tech. grade, 90%), trioctylphosphine (TOP, tech. grade, 90%), and 1,2ethanedithiol (EDT, 98%) were purchased from Alfa Aesar. Oleylamine 983

DOI: 10.1021/acs.chemmater.7b04825 Chem. Mater. 2018, 30, 982−989

Article

Chemistry of Materials



RESULTS AND DISCUSSION As we reported previously, the growth of PbSe QDs in the viscous PbCl2−OLA complex follows a diffusion controlled mode, facilitating the QD growth control to manage wide range size.23,27−31 In addition, the chlorine containing precursor allows in situ chloride passivation, resulting in well passivated QDs.23,30,31 However, the reactivities of widely used Se precursors, TOPSe and tributylphosphine selenide (TBPSe), are too low to promote the nucleation and growth of PbSe QDs in the PbCl2−OLA complex.25 It was found that the second phosphine impurities in TOP, such as dioctylphosphine, are responsible for the nucleation of QDs in the traditional synthesis in ODE.32 Therefore, commercially available second phosphine, DPP, was added to promote the chemical yield.33−35 Experiments and theoretical calculation revealed that in the presence of DPP, TOPSe can be converted to DPPSe, which is much more reactive than TOPSe.36,37 Inspired by this work, we combined DPPSe and TOPSe as a Se precursor with high and tunable reactivity. The DPPSe reacts with PbCl2−OLA complex even at room temperature to form extremely small PbSe QDs. The Se in TOPSe is gradually released via converting to DPPSe as the increase of temperature, maintaining a high oversaturation, which is critical for growth and size distribution of QDs.38,39 In this way, all the precursors can be combined at room temperature and the growth will be promoted by simply heating the mixture. As indicated by the evolution of absorption spectrum in Figure 1a, a sharp exciton peak emerged even at 30 °C, which was not affected without the presence of TOPSe in a control experiment. This indicates that the DPPSe is reactive enough to form monodispersive PbSe QDs even at room temperature. The first

exciton peak red-shifted continuously as the temperature increased from 30 to 160 °C. It is worth noting that the monodispersity was preserved in the growth demonstrated by the sharp first, second and third exciton peaks. The particle number remained constant in the early stage (30−70 °C), then decreased substantially from 70 to 110 °C, and kept almost constant in the rest stage of the growth (Figure 1b). The decrease of the particle number is a typical result of Ostwald ripening, which usually leads to broadening of size distribution. The PbSe QDs grew in a diffusion controlled mode due to the high viscosity of the Pb precursor (PbCl2−OLA complex),31 and the dissolved particles released a large amount of Se precursor maintaining the oversaturation, as a result the size distribution remained narrow, which is consistent with previous observation.30,31 Additionally, the decrease of particle number (70−110 °C) indicates the conversion of TOPSe to DPPSe in the presence of DPP is temperature dependent, that is, the reaction did not occur at low temperatures (2000 nm) thin film also shows outstanding stability in air demonstrated by the unchanged absorption spectrum in 7 days in air (Figure 5c). The stable electronic coupling PbSe QD thin films would substantially facilitate its applications in electronic and optoelectronic devices. We attribute the high stability of these PbSe QDs due to Cl− passivation. Theoretical and experimental studies have shown that halide repairs the under-coordinated sites on the surface of lead chalcogenide QDs, rendering perfect hybrid surface passivation.21,30,31,46 Because the chloride, which comes from PbCl2, is present in the reaction solution during nucleation and growth of QDs, surface reconstruction might allow for optimal surface passivation. Therefore, the in situ chloride passivation can result in better effect than those postsynthetic chloride treatments. For example, the EDT treated PbSe QD thin films here are stable in air while those made of PbSe QDs obtained by postsynthetic halide salt treatment degrade gradually in air (continual redshift of the firstexciton peak).21 The Cl− content in the PbSe QDs synthesized by our new method is measured to be 42% (compared to Pb) by XPS, which is consistent with previous results. There is no change after the PbSe QD powder was heated at 80 °C for 12 h in air as shown in Figure 6, that is, no oxidized species was formed in the heating process, which accounts for the

Figure 4. Absorption spectra of a large range of size of PbSe QDs and three typical TEM images.

size will show strong quantum confinement and might exhibit new optoelectronic properties. As a heating up method with controllability to release Se precursor, the new synthesis is capable of large scale production. By scaling up the amounts of precursors, we have managed to synthesize 23.5 g PbSe QDs from one batch of reaction as shown in Figure S3. It is worthy to note that the monodispersity was maintained in the large scale synthesis demonstrated by the sharp first, second and third exciton peaks. The as-synthesized PbSe QDs show excellent air stability in term of optical property. For lead chalcogenide QDs, the surface is exposed with (111), (110), and (100) crystalline faces.42 The

Figure 5. (a) Absorption spectra of fresh and heated (80 °C in air for 12 h in the form of powder) PbSe QDs with four different sizes. (b) Evolutions of PL QYs of PbSe QDs with first exciton peak at 1230 and 1484 nm when heated at 80 °C. (c) Absorption spectra of fresh and aged EDT treated PbSe QD thin film. 986

DOI: 10.1021/acs.chemmater.7b04825 Chem. Mater. 2018, 30, 982−989

Article

Chemistry of Materials

electronic coupled PbSe QD thin films also showed excellent stability in air, facilitating their applications in different devices. Extremely slow Ostwald ripening was observed in the new synthesis, facilitating the accurate size control of the PbSe QDs. The growth was carefully controlled to produce a large range of sizes of PbSe QDs with the first exciton peak from 750 nm to above 2100 nm. As a heating up method, this new synthesis was easily scaled up to produce 23.5 g PbSe QDs from one batch of reaction. Monodispersive CdSe and ZnSe QDs were also synthesized using this new strategy, which might be efficient for other metal selenide QDs.



Figure 6. XPS spectra of fresh and heated PbSe QDs (Pb-X = Pb-oleate, Pb−Cl, PbO, Pb(OH)2).

ASSOCIATED CONTENT

S Supporting Information *

highly stable absorption spectra and PL QYs. On the contrary, the XPS spectra changed due to the increased contribution of PbO, Pb(OH)2, SeO2, SeO32− when the PbSe QDs synthesized by conventional method was stored in air in hexane.18 Finally, to demonstrate the versatility of the strategy of in situ tuning the reactivity of Se precursor, we extended the synthetic strategy to CdSe and ZnSe QDs. In both syntheses, the size distributions became better as the increase of temperature, demonstrated by the increasing sharpness of the exciton peaks as shown in Figure 7. Injection of additional Se precursor was also

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.chemmater.7b04825. Absorption spectra of PbSe and CdSe QDs; typical TEM image and XRD pattern of PbSe QDs; absorption spectrum and picture of 23.5 g PbSe QDs and 3.87 g CdSe QDs (PDF)



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. ORCID

Daoli Zhang: 0000-0003-0646-1572 Kai Wang: 0000-0003-0443-6955 Jianbing Zhang: 0000-0003-0642-3939 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This work was supported by the National Natural Science Foundation of China (NSFC Grant Nos. 51302096, 61571206, 51402148), the Hubei Provincial Natural Science Foundation of China (ZRMS2017000370), the Fundamental Research Funds for the Central Universities (Nos. 2017KFYXJJ039, 2015TS051), the Innovation Foundation of Shenzhen Government (JCYJ20160429182959405), and the Fundamental Research Funds of Wuhan City (No. 2016060101010075). The authors thank the Analytical and Testing Center of Huazhong University of Science and Technology for the help on measurements.

Figure 7. Evolution of absorption spectrum in the synthesis of CdSe and ZnSe QDs via in situ tuning the reactivity of Se precursor.

adopted to achieve a large range of size for CdSe QDs (Figure S4). Using the heating up method, the synthesis was easily scaled up to produce 3.87 g monodispersive CdSe QDs as shown in Figure S5.





CONCLUSION A strategy of in situ tuning the reactivity of Se precursor was developed and was successfully adopted to produce ultrastable PbSe QDs. DPPSe and TOPSe were both used as the Se precursor. The reactive DPPSe induced the formation of small nanoparticles while the TOPSe did not take part in the reaction at room temperature. As the increase of the solution temperature, the DPP (in DPPSe) was converted TOPSe to reactive DPPSe, promoting the growth of the PbSe QDs and maintaining a high oversaturation condition. PbCl2−oleylamine complex was used as the Pb precursor, which allowed in situ chloride passivation. As a result, the as prepared PbSe QDs showed excellent stability in air and high PL QYs for different sizes. The first exciton peaks remained unchanged when the PbSe QDs powder were heated at 80 °C in air for 12 h. Furthermore, the ligand exchanged

REFERENCES

(1) Kagan, C. R.; Lifshitz, E.; Sargent, E. H.; Talapin, D. V. Building devices from colloidal quantum dots. Science 2016, 353 (6302), aac5523. (2) Kovalenko, M. V.; Manna, L.; Cabot, A.; Hens, Z.; Talapin, D. V.; Kagan, C. R.; Klimov, V. I.; Rogach, A. L.; Reiss, P.; Milliron, D. J.; Guyot-Sionnnest, P.; Konstantatos, G.; Parak, W. J.; Hyeon, T.; Korgel, B. A.; Murray, C. B.; Heiss, W. Prospects of Nanoscience with Nanocrystals. ACS Nano 2015, 9 (2), 1012−1057. (3) Beard, M. C. Multiple Exciton Generation in Semiconductor Quantum Dots. J. Phys. Chem. Lett. 2011, 2 (11), 1282−1288. (4) Adinolfi, V.; Sargent, E. H. Photovoltage field-effect transistors. Nature 2017, 542, 324−327. (5) Liu, M.; Voznyy, O.; Sabatini, R.; Garcia de Arquer, F. P.; Munir, R.; Balawi, A. H.; Lan, X.; Fan, F.; Walters, G.; Kirmani, A. R.; Hoogland, S.; Laquai, F.; Amassian, A.; Sargent, E. H. Hybrid organic-inorganic inks flatten the energy landscape in colloidal quantum dot solids. Nat. Mater. 2017, 16 (2), 258−263.

987

DOI: 10.1021/acs.chemmater.7b04825 Chem. Mater. 2018, 30, 982−989

Article

Chemistry of Materials (6) Wise, F. W. Lead Salt Quantum Dots: the Limit of Strong Quantum Confinement. Acc. Chem. Res. 2000, 33 (11), 773−780. (7) Shabaev, A.; Efros, A. L.; Efros, A. L. Dark and Photo-Conductivity in Ordered Array of Nanocrystals. Nano Lett. 2013, 13 (11), 5454− 5461. (8) Midgett, A. G.; Luther, J. M.; Stewart, J. T.; Smith, D. K.; Padilha, L. A.; Klimov, V. I.; Nozik, A. J.; Beard, M. C. Size and Composition Dependent Multiple Exciton Generation Efficiency in PbS, PbSe, and PbSSe Alloyed Quantum Dots. Nano Lett. 2013, 13, 3078−3085. (9) Semonin, O. E.; Luther, J. M.; Choi, S.; Chen, H. Y.; Gao, J. B.; Nozik, A. J.; Beard, M. C. Peak External Photocurrent Quantum Efficiency Exceeding 100% via MEG in a Quantum Dot Solar Cell. Science 2011, 334 (6062), 1530−1533. (10) Davis, N. J.; Bohm, M. L.; Tabachnyk, M.; Wisnivesky-RoccaRivarola, F.; Jellicoe, T. C.; Ducati, C.; Ehrler, B.; Greenham, N. C. Multiple-exciton generation in lead selenide nanorod solar cells with external quantum efficiencies exceeding 120%. Nat. Commun. 2015, 6, 8259. (11) Ehrler, B.; Musselman, K. P.; Bohm, M. L.; Morgenstern, F. S.; Vaynzof, Y.; Walker, B. J.; Macmanus-Driscoll, J. L.; Greenham, N. C. Preventing interfacial recombination in colloidal quantum dot solar cells by doping the metal oxide. ACS Nano 2013, 7 (5), 4210−4220. (12) Talgorn, E.; Gao, Y. N.; Aerts, M.; Kunneman, L. T.; Schins, J. M.; Savenije, T. J.; van Huis, M. A.; van der Zant, H. S. J.; Houtepen, A. J.; Siebbeles, L. D. A. Unity quantum yield of photogenerated charges and band-like transport in quantum-dot solids. Nat. Nanotechnol. 2011, 6 (11), 733−739. (13) Whitham, K.; Yang, J.; Savitzky, B. H.; Kourkoutis, L. F.; Wise, F.; Hanrath, T. Charge transport and localization in atomically coherent quantum dot solids. Nat. Mater. 2016, 15 (5), 557−563. (14) Delerue, C. Nanocrystal solids: Order and progress. Nat. Mater. 2016, 15 (5), 498−499. (15) Evers, W. H.; Schins, J. M.; Aerts, M.; Kulkarni, A.; Capiod, P.; Berthe, M.; Grandidier, B.; Delerue, C.; van der Zant, H. S.; van Overbeek, C.; Peters, J. L.; Vanmaekelbergh, D.; Siebbeles, L. D. High charge mobility in two-dimensional percolative networks of PbSe quantum dots connected by atomic bonds. Nat. Commun. 2015, 6, 8195. (16) Sandeep, C. S. S.; Azpiroz, J. M.; Evers, W. H.; Boehme, S. C.; Moreels, I.; Kinge, S.; Siebbeles, L. D. A.; Infante, I.; Houtepen, A. J. Epitaxially Connected PbSe Quantum-Dot Films: Controlled Neck Formation and Optoelectronic Properties. ACS Nano 2014, 8 (11), 11499−11511. (17) Dai, Q.; Wang, Y.; Zhang, Y.; Li, X.; Li, R.; Zou, B.; Seo, J.; Wang, Y.; Liu, M.; Yu, W. W. Stability Study of PbSe Semiconductor Nanocrystals over Concentration, Size, Atmosphere, and Light Exposure. Langmuir 2009, 25 (20), 12320−12324. (18) Sykora, M.; Koposov, A. Y.; McGuire, J. A.; Schulze, R. K.; Tretiak, O.; Pietryga, J. M.; Klimov, V. I. Effect of Air Exposure on Surface Properties, Electronic Structure, and Carrier Relaxation in PbSe Nanocrystals. ACS Nano 2010, 4 (4), 2021−2034. (19) Pietryga, J. M.; Werder, D. J.; Williams, D. J.; Casson, J. L.; Schaller, R. D.; Klimov, V. I.; Hollingsworth, J. A. Utilizing the Lability of Lead Selenide to Produce Heterostructured Nanocrystals with Bright, Stable Infrared Emission. J. Am. Chem. Soc. 2008, 130 (14), 4879−4885. (20) Bae, W. K.; Joo, J.; Padilha, L. A.; Won, J.; Lee, D. C.; Lin, Q.; Koh, W. K.; Luo, H.; Klimov, V. I.; Pietryga, J. M. Highly effective surface passivation of PbSe quantum dots through reaction with molecular chlorine. J. Am. Chem. Soc. 2012, 134 (49), 20160−20168. (21) Woo, J. Y.; Ko, J.-H.; Song, J. H.; Kim, K.; Choi, H.; Kim, Y.-H.; Lee, D. C.; Jeong, S. Ultrastable PbSe Nanocrystal Quantum Dots via in Situ Formation of Atomically Thin Halide Adlayers on PbSe(100). J. Am. Chem. Soc. 2014, 136 (25), 8883−8886. (22) Woo, J. Y.; Lee, S.; Lee, S.; Kim, W. D.; Lee, K.; Kim, K.; An, H. J.; Lee, D. C.; Jeong, S. Air-Stable PbSe Nanocrystals Passivated by Phosphonic Acids. J. Am. Chem. Soc. 2016, 138 (3), 876−883. (23) Zhang, J.; Gao, J.; Church, C. P.; Miller, E. M.; Luther, J. M.; Klimov, V. I.; Beard, M. C. PbSe Quantum Dot Solar Cells with More than 6% Efficiency Fabricated in Ambient Atmosphere. Nano Lett. 2014, 14 (10), 6010−6015.

(24) Zhang, J.; Chernomordik, B. D.; Crisp, R. W.; Kroupa, D. M.; Luther, J. M.; Miller, E. M.; Gao, J.; Beard, M. C. Preparation of Cd/Pb Chalcogenide Heterostructured Janus Particles via Controllable Cation Exchange. ACS Nano 2015, 9 (7), 7151−7163. (25) Zhang, C.; Xia, Y.; Zhang, Z.; Huang, Z.; Lian, L.; Miao, X.; Zhang, D.; Beard, M. C.; Zhang, J. Combination of Cation Exchange and Quantized Ostwald Ripening for Controlling Size Distribution of Lead Chalcogenide Quantum Dots. Chem. Mater. 2017, 29 (8), 3615−3622. (26) Dai, Q.; Wang, Y.; Li, X.; Zhang, Y.; Pellegrino, D. J.; Zhao, M.; Zou, B.; Seo, J.; Wang, Y.; Yu, W. W. Size-Dependent Composition and Molar Extinction Coefficient of PbSe Semiconductor Nanocrystals. ACS Nano 2009, 3 (6), 1518−1524. (27) Joo, J.; Na, H. B.; Yu, T.; Yu, J. H.; Kim, Y. W.; Wu, F.; Zhang, J. Z.; Hyeon, T. Generalized and Facile Synthesis of Semiconducting Metal Sulfide Nanocrystals. J. Am. Chem. Soc. 2003, 125 (36), 11100−11105. (28) Cademartiri, L.; Bertolotti, J.; Sapienza, R.; Wiersma, D. S.; von Freymann, G.; Ozin, G. A. Multigram Scale, Solventless, and DiffusionControlled Route to Highly Monodisperse PbS Nanocrystals. J. Phys. Chem. B 2006, 110 (2), 671−673. (29) Moreels, I.; Lambert, K.; Smeets, D.; De Muynck, D.; Nollet, T.; Martins, J. C.; Vanhaecke, F.; Vantomme, A.; Delerue, C.; Allan, G.; Hens, Z. Size-Dependent Optical Properties of Colloidal PbS Quantum Dots. ACS Nano 2009, 3 (10), 3023−3030. (30) Weidman, M. C.; Beck, M. E.; Hoffman, R. S.; Prins, F.; Tisdale, W. A. Monodisperse, Air-Stable PbS Nanocrystals via Precursor Stoichiometry Control. ACS Nano 2014, 8 (6), 6363−6371. (31) Zhang, J.; Gao, J.; Miller, E. M.; Luther, J. M.; Beard, M. C. Diffusion-Controlled Synthesis of PbS and PbSe Quantum Dots with in Situ Halide Passivation for Quantum Dot Solar Cells. ACS Nano 2014, 8 (1), 614−622. (32) Evans, C. M.; Evans, M. E.; Krauss, T. D. Mysteries of TOPSe Revealed: Insights into Quantum Dot Nucleation. J. Am. Chem. Soc. 2010, 132 (32), 10973−10975. (33) Ouyang, J. Y.; Schuurmans, C.; Zhang, Y. G.; Nagelkerke, R.; Wu, X. H.; Kingston, D.; Wang, Z. Y.; Wilkinson, D.; Li, C. S.; Leek, D. M.; Tao, Y.; Yu, K. Low-Temperature Approach to High-Yield and Reproducible Syntheses of High-Quality Small-Sized PbSe Colloidal Nanocrystals for Photovoltaic Applications. ACS Appl. Mater. Interfaces 2011, 3 (2), 553−565. (34) Joo, J.; Pietryga, J. M.; McGuire, J. A.; Jeon, S. H.; Williams, D. J.; Wang, H. L.; Klimov, V. I. A Reduction Pathway in the Synthesis of PbSe Nanocrystal Quantum Dots. J. Am. Chem. Soc. 2009, 131 (30), 10620− 10628. (35) Yu, K.; Ouyang, J. Y.; Leek, D. M. In-Situ Observation of Nucleation and Growth of PbSe Magic-Sized Nanoclusters and Regular Nanocrystals. Small 2011, 7 (15), 2250−2262. (36) Yu, K.; Liu, X.; Zeng, Q.; Leek, D. M.; Ouyang, J.; Whitmore, K. M.; Ripmeester, J. A.; Tao, Y.; Yang, M. Effect of Tertiary and Secondary Phosphines on Low-Temperature Formation of Quantum Dots. Angew. Chem., Int. Ed. 2013, 52 (18), 4823−4828. (37) Ruberu, T. P. A.; Albright, H. R.; Callis, B.; Ward, B.; Cisneros, J.; Fan, H. J.; Vela, J. Molecular Control of the Nanoscale: Effect of Phosphine-Chalcogenide Reactivity on CdS-CdSe Nanocrystal Composition and Morphology. ACS Nano 2012, 6 (6), 5348−5359. (38) Peng, X.; Wickham, J.; Alivisatos, A. P. Kinetics of II-VI and III-V colloidal semiconductor nanocrystal growth: ’focusing’ of size distributions. J. Am. Chem. Soc. 1998, 120 (21), 5343−5344. (39) Talapin, D. V.; Rogach, A. L.; Haase, M.; Weller, H. Evolution of an Ensemble of Nanoparticles in a Colloidal Solution: Theoretical Study. J. Phys. Chem. B 2001, 105 (49), 12278−12285. (40) Hendricks, M. P.; Campos, M. P.; Cleveland, G. T.; Jen-La Plante, I.; Owen, J. S. A tunable library of substituted thiourea precursors to metal sulfide nanocrystals. Science 2015, 348 (6240), 1226−1230. (41) Liu, X.; Liu, Y.; Xu, S.; Geng, C.; Xie, Y.; Zhang, Z.-H.; Zhang, Y.; Bi, W. Formation of “Steady Size” State for Accurate Size Control of CdSe and CdS Quantum Dots. J. Phys. Chem. Lett. 2017, 8, 3576−3580. (42) Schapotschnikow, P.; van Huis, M. A.; Zandbergen, H. W.; Vanmaekelbergh, D. l.; Vlugt, T. J. H. Morphological Transformations 988

DOI: 10.1021/acs.chemmater.7b04825 Chem. Mater. 2018, 30, 982−989

Article

Chemistry of Materials and Fusion of PbSe Nanocrystals Studied Using Atomistic Simulations. Nano Lett. 2010, 10 (10), 3966−3971. (43) Choi, H.; Ko, J. H.; Kim, Y. H.; Jeong, S. Steric-hindrance-driven shape transition in PbS quantum dots: understanding size-dependent stability. J. Am. Chem. Soc. 2013, 135 (14), 5278−81. (44) Moreels, I.; Fritzinger, B.; Martins, J. C.; Hens, Z. Surface chemistry of colloidal PbSe nanocrystals. J. Am. Chem. Soc. 2008, 130 (45), 15081−15086. (45) Franke, D.; Harris, D. K.; Chen, O.; Bruns, O. T.; Carr, J. A.; Wilson, M. W. B.; Bawendi, M. G. Continuous injection synthesis of indium arsenide quantum dots emissive in the short-wavelength infrared. Nat. Commun. 2016, 7, 12749. (46) Ip, A. H.; Thon, S. M.; Hoogland, S.; Voznyy, O.; Zhitomirsky, D.; Debnath, R.; Levina, L.; Rollny, L. R.; Carey, G. H.; Fischer, A.; Kemp, K. W.; Kramer, I. J.; Ning, Z.; Labelle, A. J.; Chou, K. W.; Amassian, A.; Sargent, E. H. Hybrid passivated colloidal quantum dot solids. Nat. Nanotechnol. 2012, 7 (9), 577−582.

989

DOI: 10.1021/acs.chemmater.7b04825 Chem. Mater. 2018, 30, 982−989