CsPbBr3 Solar Cells: Controlled Film Growth ... - ACS Publications

Oct 25, 2017 - postdeposition chemical treatment.26,35 CsPbBr3 QDs have also been utilized to make perovskite films for photovoltaic applications. Whi...
2 downloads 14 Views 1MB Size
Subscriber access provided by READING UNIV

Article 3

CsPbBr Solar Cells: Controlled Film Growth through Layer-by-Layer Quantum Dot Deposition Jacob B Hoffman, Gary Zaiats, Isaac Wappes, and Prashant V. Kamat Chem. Mater., Just Accepted Manuscript • DOI: 10.1021/acs.chemmater.7b03751 • Publication Date (Web): 25 Oct 2017 Downloaded from http://pubs.acs.org on October 25, 2017

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Chemistry of Materials is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 23

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

CsPbBr3 Solar Cells: Controlled Film Growth through Layer-by-Layer Quantum Dot Deposition Jacob B. Hoffman†‡, Gary Zaiats†, Isaac Wappes†‡, Prashant V. Kamat†‡§ †

Radiation Laboratory, ‡Department of Chemistry & Biochemistry, and §Department of Chemical and Biomolecular Engineering, University of Notre Dame, Notre Dame, Indiana 46556, United States ______________________________________________________________________________

* Corresponding author: Prashant V. Kamat ([email protected])

1 ACS Paragon Plus Environment

Chemistry of Materials

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Abstract: All inorganic cesium lead bromide (CsPbBr3) perovskite is a more stable alternative to methylammonium lead bromide (MAPbBr3) for designing high open circuit voltage solar cells and display devices. Poor solubility of CsBr in organic solvents make typical solution deposition methods difficult to adapt for constructing CsPbBr3 devices. Our layer-by-layer methodology, which makes use of CsPbBr3 quantum dot (QD) deposition followed by annealing, provides a convenient way to cast stable films of desired thickness. The transformation from QDs into bulk during thermal annealing arises from the resumption of nanoparticle growth and not from sintering as generally assumed. Additionally, a large loss of organic material during the annealing process is mainly from 1-octadecene left during the QD synthesis. Utilizing this deposition approach for perovskite photovoltaics is examined using typical planar architecture devices. Devices optimized to both QD spin casting concentration and overall CsPbBr3 thickness produces champion devices that reach power conversion efficiencies of 5.5% with Voc of 1.4 V. The layered QD deposition demonstrates a controlled perovskite film architecture for developing efficient, high open circuit photovoltaic devices.

TOC graphic:

2 ACS Paragon Plus Environment

Page 2 of 23

Page 3 of 23

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

Introduction Lead halide (APbX3, X = Cl-, Br-, I-) perovskite semiconductors have emerged as an attractive material for light harvesting applications,1–7 possessing superior properties for charge separation coupled with simple solution processability.8–17 Recent interest in this class of materials for photovoltaics has led to a rapid rise in the performance of laboratory assembled devices, with a highest reported power conversion efficiency of 22.1% for methylammonium (MA) lead iodide.6 The significant interest in perovskite structure semiconductors has also led to the development of perovskite structure nanomaterials. Such materials can be easily spin or drop cast on substrates, forming well packed, ordered films.18–20 Once deposited, APbX3 nanomaterials have been shown to have high PL quantum yields and low thresholds for amplified spontaneous emission, making such films promising candidates for lighting and lasing applications.18–24 Films of perovskite structured nanomaterials have also been under investigation for photovoltaic applications in CsPbX3 perovskites.25–27 The solution based deposition methods typically used for MAPbX3 are difficult to directly adapt for CsPbX3, due to lower solubility of cesium precursors in commonly used deposition solvents.28–30 Additionally, the cesium perovskite possessing the smallest band gap, CsPbI3, is not structurally stable as a bulk crystal.31–34 CsPbI3 quantum dots (QDs) have been utilized to overcome both difficulties through direct deposition of CsPbI3 QDs by spin-casting followed with stabilization of the perovskite structure through post deposition chemical treatment.26,35 CsPbBr3 QDs have also been utilized to make perovskite films for photovoltaic applications. While overall PCE is limited by a relatively larger bandgap,36 CsPbBr3 has been targeted as a potential material for stable high voltage solar cells.27–29 Specifically, Manna and co-workers have synthesized semi-suspended CsPbBr3 nanocrystal “inks” 3 ACS Paragon Plus Environment

Chemistry of Materials

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

that feature an orthorhombic perovskite structure (γ-phase). Deposition through this route results in good surface coverage and devices possessing high open circuit voltages.27 Recently, we have developed a layer-by-layer deposition procedure for bulk CsPbBr3 from fully suspended CsPbBr3 QDs that provides controllable thickness and a cubic perovskite crystal structure.37 In this report, the growth process of deposited CsPbBr3 QDs during film formation is more closely examined through electron microscopy and thermogravimetric analysis. The mechanism for the QD to bulk transformation and the performance of photovoltaic devices constructed though the layer-by-layer deposition method are discussed. Experimental Details Materials Acetone (HPLC grade, Fischer Scientific), ethanol (200 proof, anhydrous, Koptec), n-heptane (99%, spectrophotometric grade, Sigma Aldrich), n-hexane (95%, anhydrous, Sigma Aldrich), cesium carbonate (Cs2CO3, 99.9%, metals basis, Alfa Aesar), lead (II) bromide (PbBr2, 98%, Sigma Aldrich), 1-octadecene (1-ODE) (95%, Sigma Aldrich), oleic acid (OAc) (technical grade, 90%, Sigma Aldrich), oleylamine (OAm) (technical grade, 70%, Sigma Aldrich), chlorobenzene (Alfa Aesar, 99.5%), FK 102 Co(III) TFSI salt (Co[PyPz]3[TFSI]3, Dyesol), lithium bistrifluoromethanesulfonimide (LiTFSI, Sigma-Aldrich, 99.95%), and titanium diisopropoxide bis(acetylacetonate) (TAA, Sigma-Aldrich, 75 wt % in IPA), and fluorine doped tin oxide conducting glass (FTO, TECTM 7, Pilkington Glass), were used without further purification. Synthesis and Cleaning of Cesium Lead Bromide Quantum Dots The synthesis of CsPbBr3 QDs followed a procedure published by Kovalenko et al. with alterations to scale up the reaction.21 A precursor solution of cesium oleate (Cs-oleate) was 4 ACS Paragon Plus Environment

Page 4 of 23

Page 5 of 23

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

prepared in a 25 ml three neck flask by suspending Cs2CO3 (0.154 g) in OAc (1.5 ml) and 1-ODE (1.4 ml). The suspension was heated to 100 °C and placed under vacuum for 1 hour. During this time Cs2CO3 fully dissolved, forming Cs-oleate. After degassing the solution was put under nitrogen and heated to 150 °C. In a second 25 ml 3-neck flask PbBr2 (0.414 g) was suspended in OAc (3 ml) and 1-ODE (3 ml), and then was degassed under vacuum at 100 °C for 1 hour. The reaction mixture was put under nitrogen atmosphere and heated to 170 °C where OAm (3 ml) was added. After PbBr2 was fully dissolved and the solution had recovered to 170 °C, 2 ml of 150 °C Cs-oleate precursor was injected. CsPbBr3 QDs immediately formed and precipitated. Then the reaction was cooled down to 90 °C using an ice water bath. After cooling, the product was divided among four centrifuge tubes with 40 ml of 1-ODE split between tubes to prevent OAc and OAm from freezing. The precipitated QDs were separated from the reaction mixture via centrifugation at 7800 rpm for 10 minutes. Excess ligand was removed by rinsing each centrifuge tube with 5 ml 1-ODE. Excess 1-ODE was discarded followed by each tube being rinsed with acetone and dried under compressed air. QDs were then re-suspended in a solution of 9:1 (by volume) hexane:heptane and centrifuged again to remove any non-suspended material. Deposition of Cesium Lead Bromide Films Prior to deposition FTO conducting glass was cleaned through sonication in ethanol for 5 minute followed by oxygen plasma cleaning for 5 minute. Films of bulk CsPbBr3 were deposited using a modified layer-by-layer deposition procedure previously developed by our group.37 Per cycle, 15 μl of CsPbBr3 QDs were deposited through dynamic spin casting at 5,000 rpm for 30 seconds on FTO using a Hamilton syringe. The concentration of the deposition solution was standardized to produce the desired deposition thickness per cycle. After QD deposition, films 5 ACS Paragon Plus Environment

Chemistry of Materials

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

were annealed in a fume hood for 3 minutes on a hot plate set to 225 °C. At high deposition solution concentrations, vapor was observed to rise from the film during the annealing stage. Directly following annealing, the films were removed from the hot plate and allowed to cool for 1 minute. After cooling, films were dipped into hexane in attempt to remove residual 1-ODE and capping ligands. The procedure was then repeated to achieve the desired film thickness. It is important to use personal protective equipment (PPE) while carrying out above mentioned procedures. All our CsPbBr3 film deposition was conducted in fume hood with controlled ventilation and exhaust to exercise safety in handling lead compounds and volatile organics. Microscopy Experiments In situ temperature controlled transmission electron microscopy (TEM) images of annealed CsPbBr3 QDs were obtained with an FEI Titan 80-300 microscope operated at 300 keV and equipped with the DENS Solutions Inc. “wildfire system”. QDs were rinsed with acetonitrile prior to experiments to remove excess organics. Scanning electron microscopy (SEM) images were captured using an FEI-Magellan 400 scanning electron microscope with a Schottky field emitter source mounted on the gun module. Thermogravimetric and Differential Scan Calorimetry Experiments Thermogravimetric analysis (TGA) and differential scan calorimetry (DSC) experiments were performed using a Mettler-Toledo TGA/DSC 1 STARe system. TGA and DSC were run simultaneously at a scan rate of 10 °C/min under both air and nitrogen atmospheres. An aluminum crucible was used for the various organics tested with a temperature range from 28-500 °C. QD samples were dried on a hot plate set to 60 °C to remove hexane prior to the run. An alumina crucible was used in the QD experiment to scan a greater temperature range (28-900°C). 6 ACS Paragon Plus Environment

Page 6 of 23

Page 7 of 23

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

Device Assembly and Characterization An FTO substrate was masked and the exposed area was etched by reacting zinc powder with 2 M hydrochloric acid for 5 minutes. After thorough rinsing of the etched area with deionized water, the substrates were cleaned through sonication in an ethanol bath for 15 minutes, followed by an oxygen plasma treatment for 5 minutes. A TiO2 layer was deposited through spin casting a 0.3 M solution of titanium diisopropoxide bis-(acetylacetonate) in anhydrous 1-butanol at 3000 rpm for 30 seconds. The spin cast films were dried on a hot plate at 120 °C for 10 minutes before annealing at 500 °C for 30 minutes in a high temperature oven. The CsPbBr3 perovskite layer was deposited to the desired film thickness using the procedure described in the above section. Spiro-OMeTAD was deposited as a hole transport layer via spin casting at 3000 rpm in a nitrogen glove box with water content less than 5 ppm. The solution was composed of 72.3 mg/ml spiro-oMeTAD, 28.8 μl/ml 4-tert-butylpyridine, 17.5 μl/ml of a bis(trifluoromethane) sulfonimide lithium salt solution, and 10 mol% (relative to spiro-oMeTAD) of FK102 Co(II) PF6 salt in anhydrous chlorobenzene. The bis(trifluoromethane) sulfonimide lithium and FK102 Co(II) PF6 salts were added through solutions consisting of 520 and 300 mg/ml of each respective material in anhydrous acetonitrile. The electrodes were kept overnight in a dry air desiccator before a 100 nm thick gold counter electrode was deposited through thermal evaporation using a shadow mask at 10-6 torr at a 10 Å/s deposition rate. Finally, indium contacts were soldered on the cells and the finished devices were stored in a dry air desiccator until testing. Photocurrent voltage curves were recorded using a Princeton Applied Research 2273 (PARstat) potentiostat. As a light source, a simulated solar spectrum was generated from a 300 W xenon arc lamp equipped with an AM 1.5G filter. The power density of the light source was 100 mW/cm2, measured using a Thorlabs S302C thermal power sensor. Voltage was scanned at a rate 7 ACS Paragon Plus Environment

Chemistry of Materials

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

of 10 mV/s in both forward (Jsc to Voc) and reverse (Voc to Jsc) directions to minimize impacts from hysteresis. Prior to measurement cells were allowed to equilibrate for 30 seconds at Jsc and Voc for forward and reverse scan directions, respectively. Quantum Dot Growth Observed through Microscopy

Scheme 1: Schematic of QD layer-by-layer deposition. QDs are spin cast onto a substrate followed by thermal annealing for 1 minute at 225 ℃ to induce transformation of QDs into bulk. The process can then be repeated to create a bulk perovskite film of desired thickness.

The use of thick QDs films in photovoltaics requires replacement or removal of the long alkyl chained ligands typically used in the synthesis of colloidal QDs, so that the nanocrystals have close contact.38–40 For chalcogenide QDs, this is accomplished through either ligand exchange with short polar ligands or partial QD fusion from thermal annealing.41–48 The deposition method we developed involves a similar thermal annealing of spin cast CsPbBr3 QDs (Scheme 1).37 This step triggers a transformation of QDs into a layer of bulk crystallites, after which the procedure can be repeated to create a film of desired thickness. Our previous report showed QD growth through a loss of quantum confinement in UV-visible absorbance, as well as in SEM images post annealing that show a polycrystalline bulk film.37 While these experiments showed that a transformation to bulk occurred, it did not further our understanding of the mechanistic details of the process.

8 ACS Paragon Plus Environment

Page 8 of 23

Page 9 of 23

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

Figure 1: Top down SEM images of spincast QDs on an FTO substrate at two different magnifications for 0 (A, B), 1 (C, D), and 3 (E, F) minute annealing times. At high magnification (A, C, E), as annealing time increased, individual QDs disappeared forming larger crystals and revealing the underlying FTO substrate. At lower magnifications (B, D, F), dark non-conducting organic material deposited with the QDs gradually disappeared. Magnification was 150,000x for A, 100,000x for C and E, and 5,000x for B, D, and F. Insets in A, C, and E are histograms of CsPbBr3 crystal size.

To investigate the transformation more closely, top down SEM images of CsPbBr3 QDs were taken pre-annealing as well as at 1 and 3 minutes of annealing at 225 ℃ (Figure 1). When 9 ACS Paragon Plus Environment

Chemistry of Materials

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 23

examined at 150,000x (A) and 100,000x (C, E) magnification, top down SEM showed the transformation from QDs to bulk crystallites of CsPbBr3 perovskite. Pre-annealing (A), the QDs are clearly visible, self-aggregating into clusters as has been observed in the literature for this material.19,21 Average QD size was 16 nm (A, inset), found through TEM (Figure S1). Between clusters of QDs, darker areas of non-conducting material can be observed suggesting the presence of excess high boiling point organic material present in the spin casting solution. After 1 minute of annealing (C), CsPbBr3 QD aggregates have disappeared, forming larger crystallites with a 50 nm average size (C, inset). Additionally, areas of the underlying FTO substrate become visible. By 3 minutes (E), the CsPbBr3 crystals grow larger with an average size of 85 nm (E, inset), eliminating the smaller bulk crystallites observed at 1 minute. Our prior report involving QD deposition initially attributed this transformation to QD sintering, similar to what has been employed for chalcogenide QDs in QDSCs. However, the observation of larger crystal formation suggests that QD growth is simply resuming at the annealing temperatures. Top down SEM images were also obtained at lower magnifications (5,000x) in order to examine morphological changes on a wider scale. Before annealing (B), residual organic was clearly present on the film surface evidenced by a significantly darker image from the insulating material. After 1 minute of annealing time (D), some organic material was removed from the film surface, indicated by the presence of both dark and light surface regions. At 3 minutes of annealing time (F), the darker areas attributed to organic residue are no longer observable on the film surface. The loss organics is interesting as the annealing temperatures used are well below the ambient boiling points of the organic species present in the QD synthesis.

10 ACS Paragon Plus Environment

Page 11 of 23

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

Figure 2: In situ temperature controlled TEM images of CsPbBr3 QDs at 0 (A), 13 (B), and 30 (C) minutes (at 250 ℃). The longer growth timeframe compared to the SEM experiment is attributed to electron beam carbonization of the organic capping ligands. Larger bulk crystals are grown as QDs are consumed. Real time monitoring of CsPbBr3 growth was accomplished through in situ temperature controlled TEM images at 250 ℃, with snapshots of the sample taken every 15 seconds for 30 minutes. Figure 2 shows the progress of the growth at 0 (A), 13 (B), and (C) 30 minutes, but a compiled animation of the complete experiment is included in the supporting information. Prior to annealing, the sample showed significant aggregation on the TEM grid (A), triggered by rinsing the grid with acetonitrile. This was done to remove the excess ligand from the grid, avoiding carbonization from prolonged exposure to the instrument electron beam. After 13 minutes of annealing a significant number of the QDs disappear and a large crystallite of CsPbBr3 appears at the edge of the aggregate (B). By 30 minutes, growth has ended 11 ACS Paragon Plus Environment

Chemistry of Materials

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 23

with the majority of the QDs disappearing, the bulk crystal enlarging, and some smaller crystals remaining within the aggregate area. The growth is similar to what was observed in the SEM experiments with two major differences: the time scale of the TEM experiment was significantly longer and the majority of growth occurred at the aggregate edge. Both of these effects were attributed to an amorphous residue remaining at the initial QD positions, hindering the growth process by trapping CsPbBr3 (Figure S2). This residual organic skeleton likely arose from carbonization of the remaining ligands bound to the QD surface as other areas not directly observed show similar growth as seen SEM. The observed real time loss of smaller QDs as larger bulk crystallites enlarge confirms that thermal annealing triggers resumption of QD growth. In the absence of reaction precursors, material migrates from smaller to larger QDs in a defocusing process analogous to Ostwald ripening observed normally in hot-injection reactions.49 A similar process has been observed in films of PbS QDs when annealed and at similar temperatures.47 Tracking Organic Material Loss The loss of a significant amount of organic material through the annealing process is interesting due to the high boiling points of 1-octadecene (1-ODE) and the organic capping ligands relative to the annealing temperature. The removal of organics is necessary to achieve good performing devices as residual carbon can hamper device performance through increased resistance. During the annealing process, the loss of organic material can be observed by the naked eye as a vapor rising from the film surface. Additionally, elemental analysis through x-ray photoelectron spectroscopy (XPS) before and after annealing shows that the relative percentage of carbon drops from 79% to 50%, while relative amounts of oxygen, nitrogen, and CsPbBr3 increase (Table S1).

12 ACS Paragon Plus Environment

Page 13 of 23

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

To investigate the loss of organic material, thermogravimetric analysis (TGA) was employed under air atmosphere conditions for all individual organics present as well as with the QDs. Figure 3 (A) shows the relative mass loss versus temperature for the QDs (a) as well as 1ODE (b), oleylamine (OAm) (c), oleic acid (OAc) (d). In the QD sample, the hexane and heptane used to suspend the QDs were dried prior to the run to better simulate the conditions of a spin cast film. As temperature increased, a major mass loss was observed from 200-300 °C within the annealing temperature of 225 °C. When compared to the organics used in the CsPbBr3 QD reaction, this first and most significant mass loss closely reflects the TGA of 1-ODE (b), indicating that 1ODE is likely the major residual organic evaporating during annealing. This result is consistent with the QD synthesis and cleaning procedure, as 1-ODE is in excess during the reaction and is used to clean the QDs prior to resuspension. Above the annealing temperature a lesser secondary mass loss occurs (300-400 °C), arising from a combination of a residual tail observed in the TGA of 1-ODE (b) and the higher boiling point organic capping ligands used to stabilize the QDs (c,d). It is important to note that the organics removed during this secondary mass loss are unlikely to leave the film during the annealing stage since 300-400 °C is significantly higher than the annealing temperature. XPS supports this conclusion as a significant amount of carbon remains on the film post-annealing, and elements in OAc and OAm functional groups (oxygen and nitrogen) increase in relative percentage (Table S1). A previous report by Hodes and co-workers shows that CsPbBr3 is stable up to 600 °C, suggesting a higher annealing temperature can be used to remove the remaining organics without damage to the material.28 However, in practice the normally yellow material quickly becomes white at temperatures over 300 ℃ suggesting degradation. This degradation is supported by TGA coupled with differential scan calorimetry (DSC) of CsPbBr3 at higher temperatures (Figure S3).

13 ACS Paragon Plus Environment

Chemistry of Materials

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 23

Figure 3: TGA (A) and DSC (B) experiments at a scan rate of 10 °C/min under air for (a) CsPbBr3 QDs after solvent removal, (b) 1-ODE, (c) OAm, (d) OAc. In TGA the main QD mass loss at the annealing temperature used (225 °C) is attributed to excess 1-ODE from the QD reaction. In DSC the main QD transition matched a transition for 1-ODE at 290 °C attributed to 1-ODE combustion.

DSC (Figure 3 (B)) of the high temperature organics used in the QD reaction revealed additional information about the annealing process. Endothermic transitions corresponding to the expected ambient pressure boiling points of 1-ODE (315 °C) (b), OAm (350 °C) (c), and OAc (364 °C) (d) are observed in each respective DSC run. An additional feature is observed for both 1-ODE (b, 200-290 °C) and OAc (c, 225-322 °C), exhibiting a small exothermic transition followed by a larger endothermic transition. We attribute these features to be connected to the oxidation of each respective species as the features are not present when the experiment is ran under a nitrogen atmosphere (Figure S4). It is interesting to note the 1-ODE tail observed in TGA is not present under nitrogen atmosphere, suggesting the tail is the result of 1-ODE oxidation products. The features attributed to oxidation are present are present in the DSC scan for the QDs (a), indicating that organic oxidation is a present process occurring during film annealing.

14 ACS Paragon Plus Environment

Page 15 of 23

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

Quantum Dot Growth and Solid State Photovoltaics Our CsPbBr3 deposition method provides a simple pathway to achieve films of controllable thickness and growth rate. These features provide a potential advantage over solution based deposition methods for cesium based perovskites, which suffer from poor solubility of cesium halide salts in typical solvents used for perovskite deposition. To test the light harvesting applications of our deposition method, we constructed devices with a planer architecture using a 30 nm thick TiO2 compact layer. The benchmark materials of Spiro-oMeTAD and gold were used to complete the solar cell assembly (Figure 4A).

Figure 4: (A) Cross sectional SEM of CsPbBr3 devices (350 nm thickness) made with varying QD solution concentration. Devices that use less layers (higher QD concentration) contain more morphological defects than those with more layers (lower QD concentration). (B) Device performance as a function of CsPbBr3 deposition cycles to reach a total thickness of 350 nm for (i) PCE, (ii) Voc, (iii) Jsc, and (iv) FF. Current density-voltage curves were recorded at a scan rate of 10 mV/s forward and reverse directions. Each point represents 12-15 contacts from the same QD synthesis.

The concentration of the QD spin casting solution impacts the number of deposition cycles required to achieve an active layer of a targeted thickness, with a lower spin casting solution concentration requiring more deposition cycles. We first investigated how this parameter impacts device performance using 350 nm CsPbBr3 active layers deposited using QD solution 15 ACS Paragon Plus Environment

Chemistry of Materials

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 23

concentrations of 4.0 x 10-6 M, 8.0 x 10-6 M, and 1.2 x 10-5 M, which were found through a recently published molar extinction coefficient.50 The average efficiencies for each QD concentration are shown in Figure 4B, with each concentration represented as the number of cycles required to complete the deposition. Between solar cells with CsPbBr3 layers of 6 and 24 deposition cycles the average PCE (i) of the devices improved from 1.35% to 2.25% and 1.15% to 2.13% for the forward and reverse scan directions, respectively. The improvement in PCE came from increases in open circuit voltage (Voc) (ii), short circuit current (Jsc) (iii), and fill factor (FF) (iv), with the most significant gains seen in Jsc. We attribute the difference in performance to a more gradual deposition rate resulting in a more uniform CsPbBr3 layer. This morphological difference is shown in cross sectional SEM where CsPbBr3 layers with 6 deposition cycles contain significantly more cracks and pinholes compared to samples with 24 deposition cycles (Figure 4A). Next we investigated the effect of varied CsPbBr3 thickness on device performance. Performance parameters for devices as a function of CsPbBr3 thickness are shown in Figure 5A for both forward and reverse scan directions. The general PCE (i) of the solar cells increases with CsPbBr3 thickness until 250 nm where it levels off with an average cell performance of approximately 2%. The majority of the difference in performance comes from changes in Voc (ii), which drastically increases from 0.2 Volts at 40 nm thickness to 1.2 Volts at 300 nm thickness. Cross sectional SEM images show areas without CsPbBr3 coverage in thinner devices (Figure S5). Areas without full coverage have electron and hole transport layers in direct contact, causing cell shorting and lowering measured photovoltage. In addition to changes in Voc, Jsc (iii) gradually increased with CsPbBr3 thickness, which is attributed to thicker CsPbBr3 layers absorbing more light. This correlation is shown by direct comparison of Jsc to the light absorbed by CsPbBr3

16 ACS Paragon Plus Environment

Page 17 of 23

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

(Figure S6). Values for FF (iv) remain consistent around 0.4, suggesting active layer thickness plays a minimal role in that parameter.

Figure 5: (A) PCE (i), Voc (ii), Jsc (iii), and FF (iv) for CsPbBr3 solar cells as a function of CsPbBr3 thickness. (B) Current density-voltage curves of our champion device. Current density-voltage curves were recorded at a scan rate of 10 mV/s forward and reverse directions. Each point represents 8-15 contacts from separate batches of devices.

When compared to other devices with similar architectures, our devices have comparable values for Voc and Jsc. However, there is a difference in FF with our devices possessing lower values, leading to a lower average PCE. While the average performance of our devices is lower than those using traditional deposition methods, our champion device achieves comparable efficiencies up to 5.5% with values of 1.4 V, 7 mA/cm2, and 0.55 for Voc and Jsc, and FF, respectively (Figure 5B). In all devices, hysteresis exists as is common in perovskite type solar cells, but hysteresis is significantly diminished in our champion device. Despite comparable overall performance to typical deposition methods in champion devices, values for FF remained low. One possible explanation for low FF is the high percentage of residual carbon (50% as

17 ACS Paragon Plus Environment

Chemistry of Materials

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 23

estimated from XPS) from oxidation products and native capping ligands on the film surface post annealing, leading to increased series resistance. The comparable voltages and currents achieved with only partial organic material removal shows that perovskite deposition through QD annealing is a promising platform for device assembly. Our deposition method is similar to the efforts by Manna and co-workers featuring semisuspended CsPbBr3 nanocrystal inks synthesized at room temperature using low boiling point organics. Devices assembled this way have significantly lower relative carbon content and possess superior FF as all organics used in the synthesis can evaporate at low temperatures. However, the room temperature reaction keeps CsPbBr3 in the orthorhombic perovskite phase. Additionally, nanocrystals appear to be only semi-suspended in organic solvent, while when passivated with typical long chained capping ligands QDs are fully suspended in non-polar solvent for months after synthesis. This provides the potential advantage of long term storage prior to deposition. In the future, the best parts of both approaches could be achieved by adjusting the QD synthesis to use lower boiling point ligands that do not boil at the reaction temperature (170 ℃), but evaporate within the annealing temperature similar to 1-ODE. Conclusion In summary, an alternative method for CsPbBr3 perovskite deposition that relies on the layer-by-layer deposition and subsequent annealing of CsPbBr3 QDs was further explored as a potential way to assemble CsPbBr3 solar cells. The mechanism of the transformation from QDs to bulk was found through SEM and in situ TEM experiments. As opposed to QD sintering that was initially assumed, QD growth was found to resume at the annealing temperature with defocusing processes dominating. A loss of organic material during the film annealing was determined to be primarily from evaporation of 1-ODE, with the native capping ligands and oxidation products of 18 ACS Paragon Plus Environment

Page 19 of 23

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

1-ODE remaining post-annealing. Planar CsPbBr3 solar cells made using this deposition method were optimized by varying the QD solution concentration as well as active layer thickness, and champion devices had comparable PCE to similar devices made via typical deposition routes. Further optimization may be achieved through by tuning the QD reaction organics to evaporate during annealing. Acknowledgements: The research described herein was supported by the Division of Chemical Sciences, Geosciences, and Biosciences, Office of Basic Energy Sciences of the U.S. Department of Energy through award DE-FC02-04ER15533. This article is contribution number NDRL No. 5185 from the Notre Dame Radiation Laboratory. During the temperature controlled TEM experiments, the authors would like to acknowledge the NDIIF TEM core facility for free use of Titan TEM, DENS Solutions Inc. for providing the DENS Wildfire heating holder and nano chip, and Dr. Sergei Rouvimov for assistance during the experiment. Supporting Information: A file of sequential temperature controlled TEM images depicting QD growth can be found as additional content. The organic skeleton of the TEM experiment, a table of relative elemental composition from XPS of films pre and post annealing, an extended TGA-DSC temperature range of CsPbBr3 QDs, TGA-DSC of organics under nitrogen atmosphere, cross sectional SEM of devices using different CsPbBr3 QD concentration as well as varied CsPbBr3 thickness, and light absorbance compared to average current density at varied device thickness can be found in the supporting information. References: 19 ACS Paragon Plus Environment

Chemistry of Materials

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 20 of 23

(1)

Sum, T. C.; Mathews, N. Advancements in Perovskite Solar Cells: Photophysics behind the Photovoltaics. Energy Environ. Sci. 2014, 7 (8), 2518–2534.

(2)

Lewis, N. S. Research Opportunities to Advance Solar Energy Utilization. Science 2016, 351 (6271), 353–363.

(3)

Hodes, G. Perovskite-Based Solar Cells. Science 2013, 342 (6156), 317–318.

(4)

Heo, J. H.; Im, S. H. Highly Reproducible, Efficient Hysteresis-Less CH3NH3PbI3-xClx Planar Hybrid Solar Cells without Requiring Heat-Treatment. Nanoscale 2016, 8 (5), 2554– 2560.

(5)

Kojima, A.; Teshima, K.; Shirai, Y.; Miyasaka, T. Organometal Halide Perovskites as Visible-Light Sensitizers for Photovoltaic Cells. J. Am. Chem. Soc. 2009, 131 (17), 6050– 6051.

(6)

NREL Best Research-Cell http://www.nrel.gov/ncpv/images/efficiency_chart.jpg.

(7)

Manser, J. S.; Christians, J. A.; Kamat, P. V. Intriguing Optoelectronic Properties of Metal Halide Perovskites. Chem. Rev. 2016, 116 (21), 12956–13008.

(8)

Ponseca, C. S.; Savenije, T. J.; Abdellah, M.; Zheng, K.; Yartsev, A.; Pascher, T.; Harlang, T.; Chabera, P.; Pullerits, T.; Stepanov, A.; et al. Organometal Halide Perovskite Solar Cell Materials Rationalized: Ultrafast Charge Generation, High and Microsecond-Long Balanced Mobilities, and Slow Recombination. J. Am. Chem. Soc. 2014, 136 (14), 5189– 5192.

(9)

Yin, W.-J. W.; Yang, J.-H. J.; Kang, J.; Yan, Y.; Wei, S.-H. Halide Perovskite Materials for Solar Cells: A Theoretical Review. J. Mater. Chem. A 2015, 3 (17), 8926–8942.

(10)

Yin, W. J.; Shi, T.; Yan, Y. Unique Properties of Halide Perovskites as Possible Origins of the Superior Solar Cell Performance. Adv. Mater. 2014, 26 (27), 4653–4658.

(11)

Feng, J.; Xiao, B. Crystal Structures, Optical Properties, and Effective Mass Tensors of CH3NH3PbX3 (X = I and Br) Phases Predicted from HSE06. J. Phys. Chem. Lett. 2014, 5 (7), 1278–1282.

(12)

Frost, J. M.; Butler, K. T.; Brivio, F.; Hendon, C. H.; Van Schilfgaarde, M.; Walsh, A. Atomistic Origins of High-Performance in Hybrid Halide Perovskite Solar Cells. Nano Lett. 2014, 14 (5), 2584–2590.

(13)

Oga, H.; Saeki, A.; Ogomi, Y.; Hayase, S.; Seki, S. Improved Understanding of the Electronic and Energetic Landscapes of Perovskite Solar Cells: High Local Charge Carrier Mobility, Reduced Recombination, and Extremely Shallow Traps. J. Am. Chem. Soc. 2014, 136 (39), 13818–13825.

(14)

Stranks, S. D.; Eperon, G. E.; Grancini, G.; Menelaou, C.; Alcocer, M. J. P.; Leijtens, T.; Herz, L. M.; Petrozza, A.; Snaith, H. J. Electron-Hole Diffusion Lengths Exceeding 1 Micrometer in an Organometal Trihalide Perovskite Absrober. Science 2014, 341 (6156), 341–345.

20 ACS Paragon Plus Environment

Efficiencies

Page 21 of 23

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

(15)

Dong, Q.; Fang, Y.; Shao, Y.; Mulligan, P.; Qiu, J.; Cao, L.; Huang, J. Electron-Hole Diffusion Lengths > 175 μm in Solution-Grown CH3NH3PbI3 Single Crystals. Science 2015, 347 (6225), 967–970.

(16)

De Wolf, S.; Holovsky, J.; Moon, S. J.; Löper, P.; Niesen, B.; Ledinsky, M.; Haug, F. J.; Yum, J. H.; Ballif, C. Organometallic Halide Perovskites: Sharp Optical Absorption Edge and its Relation to Photovoltaic Performance. J. Phys. Chem. Lett. 2014, 5 (6), 1035–1039.

(17)

Xing, G.; Mathews, N.; Sun, S.; Lim, S. S.; Lam, Y. M.; Grätzel, M.; Mhaisalkar, S.; Sum, T. C. Long-Range Balanced Electron- and Hole-Transport Lengths in Organic-Inorganic CH3NH3PbI3. Science 2013, 342 (6156), 344–347.

(18)

Bekenstein, Y.; Koscher, B. A.; Eaton, S. W.; Yang, P.; Alivisatos, A. P. Highly Luminescent Colloidal Nanoplates of Perovskite Cesium Lead Halide and their Oriented Assemblies. J. Am. Chem. Soc. 2015, 137 (51), 16008–16011.

(19)

Pan, J.; Sarmah, S. P.; Murali, B.; Dursun, I.; Peng, W.; Parida, M. R.; Liu, J.; Sinatra, L.; Alyami, N.; Zhao, C.; et al. Air-Stable Surface-Passivated Perovskite Quantum Dots for Ultra-Robust, Single- and Two-Photon-Induced Amplified Spontaneous Emission. J. Phys. Chem. Lett. 2015, 6 (24), 5027–5033.

(20)

Veldhuis, S. A.; Boix, P. P.; Yantara, N.; Li, M.; Sum, T. C.; Mathews, N.; Mhaisalkar, S. G. Perovskite Materials for Light-Emitting Diodes and Lasers. Adv. Mater. 2016, 28 (32), 6804–6834.

(21)

Protesescu, L.; Yakunin, S.; Bodnarchuk, M. I.; Krieg, F.; Caputo, R.; Hendon, C. H.; Yang, R. X.; Walsh, A.; Kovalenko, M. V. Nanocrystals of Cesium Lead Halide Perovskites (CsPbX3, X = Cl, Br, and I): Novel Optoelectronic Materials Showing Bright Emission with Wide Color Gamut. Nano Lett. 2015, 15 (6), 3692–3696.

(22)

Yakunin, S.; Protesescu, L.; Krieg, F.; Bodnarchuk, M. I.; Nedelcu, G.; Humer, M.; De Luca, G.; Fiebig, M.; Heiss, W.; Kovalenko, M. V. Low-Threshold Amplified Spontaneous Emission and Lasing from Colloidal Nanocrystals of Caesium Lead Halide Perovskites. Nat. Commun. 2015, 6, 8056.

(23)

Yang, M.; Li, Z.; Reese, M. O.; Reid, O. G.; Kim, D. H.; Siol, S.; Klein, T. R.; Yan, Y.; Berry, J. J.; van Hest, M. F. A. M.; et al. Perovskite Ink with Wide Processing Window for Scalable High-Efficiency Solar Cells. Nat. Energy 2017, 2, 1–9.

(24)

Palazon, F.; Stasio, F. Di; Akkerman, Q. A.; Krahne, R.; Prato, M.; Manna, L. PolymerFree Films of Inorganic Halide Perovskite Nanocrystals as UV-to-White Color-Conversion Layers in LEDs. Chem. Mater. 2016, 28 (9), 2902−2906.

(25)

Im, J.; Luo, J.; Franckevic, M.; Pellet, N.; Gao, P.; Moehl, T.; Zakeeruddin, S. M.; Nazeeruddin, M. K.; Gra, M.; Park, N. Nanowire Perovskite Solar Cell. Nano Lett. 2015, 15 (3), 2120−2126.

(26)

Swarnkar, A.; Marshall, A. R.; Sanehira, E. M.; Chernomordik, B. D.; Moore, D. T.; Christians, J. A.; Chakrabarti, T.; Luther, J. M. Quantum Dot-Induced Phase Stabilization of α-CsPbI3 Perovskite for High-Efficiency Photovoltaics. Science 2016, 354 (6308), 92– 95. 21 ACS Paragon Plus Environment

Chemistry of Materials

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 22 of 23

(27)

Akkerman, Q. A.; Gandini, M.; Stasio, F. Di; Rastogi, P.; Palazon, F.; Bertoni, G.; Ball, J. M.; Prato, M.; Petrozza, A.; Manna, L. Strongly Emissive Perovskite Nanocrystal Inks for High-Voltage Solar Cells. Nat. Energy 2016, 2, 1–7.

(28)

Kulbak, M.; Gupta, S.; Kedem, N.; Levine, I.; Bendikov, T.; Hodes, G.; Cahen, D. Cesium Enhances Long-Term Stability of Lead Bromide Perovskite-Based Solar Cells. J. Phys. Chem. Lett. 2016, 7 (1), 167–172.

(29)

Kulbak, M.; Cahen, D.; Hodes, G. How Important Is the Organic Part of the Lead Halide Perovskite Photovoltaic Cells? Efficient CsPbBr3 Cells. J. Phys. Chem. Lett. 2015, 6 (13), 2452–2456.

(30)

Fai, C.; Lau, J.; Deng, X.; Ma, Q.; Zheng, J.; Yun, J. S.; Green, M. A.; Huang, S.; Ho-baillie, A. W. Y. CsPbIBr2 Perovskite Solar Cell by Spray-Assisted Deposition. ACS Energy Lett. 2016, 1 (3), 573–577.

(31)

Yunakova, O. N.; Miloslavskii, V. K.; Kovalenko, E. N. The Exciton Absorption Spectrum of Thin KPbI3 Thin Films. Opt. Spectrosc. 2013, 116 (1), 68–71.

(32)

Yunakova, O. N.; Miloslavskii, V. K.; Kovalenko, E. N. Exciton Absorption Spectrum of Thin CsPbI3 and Cs4PbI6 Films. Opt. Spectrosc. 2011, 112 (1), 91–96.

(33)

Eperon, G. E.; Paterno, G. M.; Sutton, R. J.; Zampetti, A.; Haghighirad, A. A.; Cacialli, F.; Snaith, H. J. Inorganic Caesium Lead Iodide Perovskite Solar Cells. J. Mater. Chem. A 2015, 3 (39), 19688–19695.

(34)

Stoumpos, C. C.; Malliakas, C. D.; Kanatzidis, M. G. Semiconducting Tin and Lead Iodide Perovskites with Organic Cations: Phase Transitions, High Mobilities, and Near-Infrared Photoluminescent Properties. Inorg. Chem. 2013, 52 (15), 9019–9038.

(35)

Dastidar, S.; Egger, D. A.; Tan, L. Z.; Cromer, S. B.; Dillon, A. D.; Liu, S.; Kronik, L.; Rappe, A. M.; Fafarman, A. T. High Chloride Doping Levels Stabilize the Perovskite Phase of Cesium Lead Iodide. Nano Lett. 2016, 16 (6), 3563–3570.

(36)

Shockley, W.; Queisser, H. J. Detailed Balance Limit of Efficiency of P-N Junction Solar Cells. J. Appl. Phys. 1961, 32 (3), 510–519.

(37)

Hoffman, J. B.; Schleper, A. L.; Kamat, P. V. Transformation of Sintered CsPbBr3 Nanocrystals to Cubic CsPbI3 and Gradient CsPbBrxI3-X through Halide Exchange. J. Am. Chem. Soc. 2016, 138 (27), 8603–8611.

(38)

Remacle, F. On Electronic Properties of Assemblies of Quantum Nanodots. J. Phys. Chem. A 2000, 104 (20), 4739–4747.

(39)

Liu, Y.; Gibbs, M.; Puthussery, J.; Gaik, S.; Ihly, R.; Hillhouse, H. W.; Law, M. Dependence of Carrier Mobility on Nanocrystal Size and Ligand Length in Pbse Nanocrystal Solids. Nano Lett. 2010, 10 (5), 1960–1969.

(40)

Carey, G. H.; Abdelhady, A. L.; Ning, Z.; Thon, S. M.; Bakr, O. M.; Sargent, E. H. Colloidal Quantum Dot Solar Cells. Chem. Rev. 2015, 115 (23), 12732–12763.

(41)

Talgorn, E.; Moysidou, E.; Abellon, R. D.; Savenije, T. J.; Goossens, A.; Houtepen, A. J.; 22 ACS Paragon Plus Environment

Page 23 of 23

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

Siebbeles, L. D. A. Highly Photoconductive CdSe Quantum-Dot Films: Influence of Capping Molecules and Film Procedure Preparation. J. Phys. Chem. C 2010, 114 (8), 3441– 3447. (42)

Ip, A. H.; Thon, S. M.; Hoogland, S.; Voznyy, O.; Zhitomirsky, D.; Debnath, R.; Levina, L.; Rollny, L. R.; Carey, G. H.; Fischer, A.; et al. Hybrid Passivated Colloidal Quantum Dot Solids. Nat. Nanotechnol. 2012, 7 (9), 577–582.

(43)

Wang, Y.; Li, X.; Song, J.; Xiao, L.; Zeng, H.; Sun, H. All-Inorganic Colloidal Perovskite Quantum Dots: A New Class of Lasing Materials with Favorable Characteristics. Adv. Mater. 2015, 27 (44), 7101–7108.

(44)

Baumgardner, W. J.; Whitham, K.; Hanrath, T. Confined-but-Connected Quantum Solids via Controlled Ligand Displacement. Nano Lett. 2013, 13 (7), 3225–3231.

(45)

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.

(46)

Tang, J.; Kemp, K. W.; Hoogland, S.; Jeong, K. S.; Liu, H.; Levina, L.; Furukawa, M.; Wang, X.; Debnath, R.; Cha, D.; et al. Colloidal-Quantum-Dot Photovoltaics Using Atomic-Ligand Passivation. Nat. Mater. 2011, 10 (10), 765–771.

(47)

Ding, B.; Wang, Y.; Huang, P.; Waldeck, D. H.; Lee, J. Depleted Bulk Heterojunctions in Thermally Annealed PbS Quantum Dot Solar Cells. J. Phys. Chem. C 2014, 118 (27), 14749–14758.

(48)

Jo, C. H.; Kim, J. H.; Kim, J.; Kim, J.; Oh, M. S.; Kang, M. S.; Kim, M.-G.; Kim, Y.-H.; Ju, B.-K.; Park, S. K. Low-Temperature Annealed PbS Quantum Dot Films for Scalable and Flexible Ambipolar Thin-Film-Transistors and Circuits. J. Mater. Chem. C 2014, 2 (48), 10305–10311.

(49)

De Smet, Y.; Deriemaeker, L.; Finsy, R. A Simple Computer Simulation of Ostwald Ripening. Langmuir 1997, 13 (26), 6884–6888.

(50)

De Roo, J.; Ibáñez, M.; Geiregat, P.; Nedelcu, G.; Walravens, W.; Maes, J.; Martins, J. C.; Van Driessche, I.; Kovalenko, M. V.; Hens, Z. Highly Dynamic Ligand Binding and Light Absorption Coefficient of Cesium Lead Bromide Perovskite Nanocrystals. ACS Nano 2016, 10 (2), 2071–2081.

23 ACS Paragon Plus Environment