Development of Low-Temperature Solution-Processed Colloidal Quantum Dot-Based Solar Cells ARCHIVES by MASSACHUSETS 1N57 OF TECHNOLOGYy Liang-Yi Chang MAY 1 4 2014 B.S. Materials Science and Engineering LIBRARIES National Tsing Hua University, 2005 Submitted to the Department of Materials Science and Engineering on August 5, 2013, in Partial Fulfillment of the Requirements for the Degree of Doctor of Philosophy in Materials Science and Engineering at the MASSACHUSETTS INSTITUTE OF TECHNOLOGY September 2013 ©Massachusetts Institute of Technology 2013. All rights reserved. .................................... Author..... Department of Materials Science and Engineering August 5, 2013 Certified by................. ............................. ......................... Moungi G. Bawendi Lester Wolfe Professor in Chemistry Thesis Supervisor ............................ Certified by........ Lionel C. Kimerling Thomas Lord Professor of Materials Science and Engineering DepartmentalTesis Co-Supervisor Accepted by..................................... Gerbrand Ceder Chair, Departmental Committee on Graduate Students f 2 Development of Low-Temperature Solution-Processed Colloidal Quantum Dot-Based Solar Cells by Liang-Yi Chang Submitted to the Department of Materials Science and Engineering on August 5, 2013, in Partial Fulfillment of the Requirements for the Degree of Doctor of Philosophy in Materials Science and Engineering Abstract Solution-processed solar cells incorporating organic semiconductors and inorganic colloidal quantum dots (QDs) are potential alternatives to conventional solar cells fabricated via vacuum or high-temperature sintering processes for large-area, high-throughput, and low-cost manufacturing. In this work, we explored two types of solution-processed QD-based solar cells: all-inorganic solar cells and organic/inorganic hybrid solar cells. In the all-inorganic device, three QD deposition techniques (spin coating, dip coating, and spray coating) were first experimented in order to prepare high-quality QD thin film for the photovoltaic application. The device was based on the heterojunction formed between dip-coated PbS QD layers and CdS thin film that was deposited via a solution process at 80'C. The resultant device, employing a 1,2-ethanedithiol ligand exchange scheme, exhibits comparable power conversion efficiency (3.5%) to that of high-temperature (260'C) sintered or sputtered ZnO/PbS (PbSe) QD devices. The initial device yield issue associated with the pinhole formation was addressed, and a procedure for the fabrication of reproducible devices was formulated. The demonstration of this device is a step towards low-cost solar cell manufacturing. Through a combination of thickness-dependent current density-voltage characteristics, optical modeling, and capacitance measurements, the combined diffusion length and depletion width in the PbS QD layer is found to be approximately 170 nm. In the organic/inorganic hybrid device, poly(3-hexylthiophene-2,5-diyl) (P3HT) nanofibers and CdS QDs were employed as electron donors and acceptors, respectively. Crystalline P3HT nanofibers, grown from amorphous P3HT solution, were blended with CdS QDs to form bulk heterojunctions. By adding a large quantity of poor solvent to the blended solution, we demonstrated preferential decoration of CdS QDs onto P3HT nanofibers and stronger interaction between these two materials. The resultant device also showed improved open-circuit voltage, short-circuit current density, and power conversion efficiency. Thesis Supervisor: Moungi G. Bawendi Title: Lester Wolfe Professor in Chemistry 3 4 Acknowledgements This thesis is the result of inspiration, collaboration, encouragement, and support from several professors, colleagues, friends, and families of mine. I am sincerely thankful for having them before, during, and after my PhD study. Prof. Tai-Bor Wu, an outstanding and esteemed professor in the Department of Materials Science and Engineering at National Tsing Hua University in Taiwan and a kind father of one of my best high school friends, introduced materials science to me and encouraged me to pursue my study in the U.S. He unfortunately passed away from a heart attack in 2010, and I would like to dedicate this thesis to him. I am grateful to have Prof. Moungi Bawendi as my thesis advisor. He gave me plenty of freedom when choosing research directions, support when equipment purchase and training was needed, and guidance and patience when I was stuck in experimental failure. He has been encouraging me to become an independent researcher. Members in the Bawendi group have been very helpful throughout my study. In particular, Dr. Scott Geyer instructed me QD synthesis and helped me to set up lab equipment for device fabrication and characterization. I also would like to thank the device subgroup, including Darcy Wanger, Gyu Weon Hwang, Jennifer Scherer, Chia-Hao Chuang, and Whitney Hess, for the discussion spanning from material synthesis and processing to device fabrication and characterization. Most importantly, we worked together to synthesize batches and batches of PbS QDs. Prof. Vladimir Bulovic and members from his group have been very generous to share the access to the Organic and Nanostructured Electronics Laboratory and to offer advise on device physics and engineering. In particular, Dr. Richard Lunt wrote the majority of the MATLab code 5 for the optical modeling carried out in this study. Patrick Brown, Dr. Tim Osedach, and Geoffrey Supran have been great resources to discuss device fabrication and characterization. Dr. Ni Zhao has always cheered me up when my experiments failed in the early days. My friends, especially Shih-Wei Chang, Chung-Hsiang Chang, Vivian Chuang, Hsu-Yi Lee, Kevin Lee, Yi-Chun Lu, Amy Chi, Carl Yu, Tsao-Hsien Chen, Yu-Chung Hsiao, Elsie Li, Chia-Hua Lee, Wei Yi, Kevin Chuang, and Kylie Yu, have undoubtedly completed my life outside the lab. We explored cities, had meals, played sports, and had a lot of fun together. The last but not least, I would like to thank my families in Taiwan for their endless love. 6 Preface Parts of this thesis were reproduced from previously published manuscripts. The associated copyright approval has been obtained as indicated below. Chapter 5 Reprinted with permission from "Liang-Yi Chang, Richard R. Lunt, Patrick R. Brown, Vladimir Bulovid, and Moungi G. Bawendi, Low-Temperature Solution-Processed Solar Cells Based on PbS Colloidal Quantum Dot/CdS Heterojunctions, Nano Letters, 2013, 13 (3), 994-999." Copyright 2013 American Chemical Society. Chapter 6 Reprinted with permission from "Shenqiang Ren, Liang-Yi Chang, Sung-Keun Lim, Jing Zhao, Matthew Smith, Ni Zhao, Vladimir Bulovid, Moungi Bawendi, and Silvija Gradebak, Inorganic-Organic Hybrid Solar Cell: Bridging Quantum Dots to Conjugated Polymer Nanowires, Nano Letters, 2011, 11 (9), 3998-4002." Copyright 2011 American Chemical Society. 7 8 Table of Contents 1 M otivation ............................................................................................................................ 22 1.1 Solar Photovoltaic Background..................................................................................... 22 1.2 Low-Temperature Solution-Processed Colloidal Quantum Dot Solar Cells ........... 24 1.3 Thesis Outline..................................................................................................................28 1.4 Reference.........................................................................................................................29 2 Basics of Colloidal Quantum Dot Solar Cells......................................................................31 2.1 Colloidal Quantum Dot Synthesis .................................................................................... 31 2.2 Electrical Properties of Colloidal Quantum Dots .............................................................. 34 2.2.1 Carrier M obility ........................................................................................................ 34 2.2.2 Carrier Type and Density ...................................................................................... 36 2.3 Solar Cell Device Physics ............................................................................................. 37 2.3.1 Photovoltaic Processes in Inorganic Solar Cells...................................................... 37 2.3.2 Photovoltaic Processes in Organic Solar Cells ....................................................... 41 2.3.3 Current-Voltage Characteristics of Solar Cells........................................................ 43 2.4 Conclusion.......................................................................................................................47 2.5 Reference.........................................................................................................................48 3 Experim ental M ethods.........................................................................................................53 3.1 Fabrication of Pbs QD/CdS Bilayer Heterojunction Devices ......................................... 53 3.1.1 PbS Colloidal QD Synthesis and Purification ....................................................... 53 9 3.1.2 Substrate Cleaning .................................................................................................... 54 3.1.3 (3-mercaptopropyl)trim ethoxysilane Substrate Treatment....................................... 54 3.1.4 Deposition of PbS QD Film ................................................................................... 55 3.1.5 Deposition of CdS Thin Film ................................................................................. 56 3.1.6 Deposition of Al Electrodes................................................................................... 59 3.2 Fabrication of P3HT Nanofiber/CdS QD Bulk Heterojunction devices..........................59 3.2.1 Growth of P3HT Nanofibers................................................................................... 59 3.2.2 CdS Colloidal QD Synthesis, Purification, and n-Butylamine Ligand Exchange ........ 60 3.2.3 Blending of P3HT N anofibers and CdS QDs ......................................................... 61 3.2.4 Deposition of PEDOT:PSS..................................................................................... 61 3.2.5 Deposition and Treatment of P3HT Nanofiber/CdS QD Layer...............................61 3.2.6 Deposition of Bathocuproine and M g:Ag/Ag Layers ............................................. 62 3.3 M aterial and Device Characterization .......................................................................... 62 3.3.1 M aterial Characterization ....................................................................................... 62 3.3.2 Device Characterization ........................................................................................ 63 3.4 Reference.........................................................................................................................63 4 PbS QD Schottky Photovoltaic Devices........................................................................... 65 4.1 Introduction ..................................................................................................................... 65 4.2 Deposition of QD Film ................................................................................................. 65 4.2.1 Spin Coating ............................................................................................................. 65 10 4.2.2 Dip Coating...............................................................................................................66 4.2.3 Spray Coating............................................................................................................66 4.3 Film M orphology and Device Characteristics...................................................................67 4.4 Conclusion.......................................................................................................................70 4.5 Reference.........................................................................................................................70 5 PbS Q D/CdS Bilayer Heterojunction Photovoltaic Devices ........................................... 72 5.1 Introduction ..................................................................................................................... 72 5.2 Device Structure and Fabrication .................................................................................. 73 5.3 Protocols for the Fabrication of Reproducible Heterojunction Devices .......................... 76 5.4 The Comparison between Schottky and PN Heterojunction Devices ............................ 83 5.5 PbS QD Layer Thickness-Dependent J-V Characteristics............................................. 84 5.6 M odeling of PbS QD Layer Thickness-Dependent Jsc ................................................. 86 5.7 Capacitance M easurement........................................................................................... 93 5.8 CdS Layer Thickness-Dependent Jsc............................................................................. 95 5.9 Conclusion.......................................................................................................................98 5.10 Reference.......................................................................................................................98 6 Efficiency Improvement in Inorganic-Organic Hybrid Photovoltaic Devices via Solvent-Assisted Self Assem bly.............................................................................................102 6.1 Introduction ................................................................................................................... 102 6.2 Device Structure and Fabrication ................................................................................... 103 11 6.3 Material Characterization: Random Mix versus Self Assembly ................... 104 6.4 Device Characterization.................................................................................................109 6.5 Conclusion..................................................................................................................... 114 6.6 Reference....................................................................................................................... 115 7 O utlook ............................................................................................................................... 118 7.1 Conclusion..................................................................................................................... 118 7.2 Future Work .................................................................................................................. 119 12 List of Tables 1-1. Price of solar electricity in US cents per kilowatt-hour reported in 2012 for various installation scales and climate conditions. The residential price is calculated based upon a standard 2 kilowatt peak system; the commercial price, a 50 kilowatt peak system; the industrial price, 500 kilowatt peak system ...................................................................... 27 1-2. Theoretical power conversion efficiency of single- and multijunction solar cells and the corresponding optimal band gap energies for the constituent junctions. Band gap energies are expressed with equivalent photon wavelength. ........................................................ 3-3. Parameters for PbS QD dip coating. A dip coating cycle consists of: (1) 30 dipping the substrate into 10mg/ml PbS QDs in hexane for 1 second and drying the sample for 30 seconds, (2) dipping the film in 0.02% EDT in acetonitrile for 30 seconds, and (3) rinsing by dipping in pure acetonitrile for 3 seconds, followed by drying for 60 seconds. The above dip coating cycle was repeated to achieve desired PbS QD thicknesses. The dipping speed for all the steps was 400 mm/min. Before and after dipping into QD solution, the samples were dried naturally for 60 and 30 seconds respectively to avoid cross-contamination between QD solution and acetonitrile or EDT solution, which can produce agglomerates in solution, resulting in poor film qualities. ............................................. 13 QD 59 14 List of Figures 1-1. Schematic of a colloidal quantum dot with inorganic semiconductor core and organic ligand sh e ll...................................................................................................................................2 8 1-2. The effect of quantum confinement on the energy levels of electrons (solid circle) and holes (open circle) in semiconductor nanoparticles. ................................................................ 29 1-3. Absorption spectra of PbS QDs with a series of particle diameters. The first exitonic absorption peak of PbS QDs can be tuned readily from 600 to 2200 nm. ....................... 30 2-1. A photograph of colloidal QD synthesis setup, including a Schlenk line, a three-neck flask, a condenser, a stiffer, a temperature controlling system, an injection gun, and a mechanical rotary pump (not shown in the picture). The black injection gun shown in this picture is used for large-scale (gram-scale) synthesis, while for a conventional small-scale synthesis (100 milligram-scale) synthesis, a syringe is used ......................................................... 36 2-2. Cartoon depicting the precursor concentration evolution along different stages of colloidal QD nucleation and growth under the framework of LaMer colloid growth model.......36 2-3. A Scanning electron microscopy image of a treated QD film. Cracks are usually developed in thick QD film after post-deposition treatment. Cracks are usually developed in thick QD film after post-deposition treatm ent ................................................................................ 2-4. Energy band diagram of a pn homojunction at thermal equilibrium. ............................... 39 43 2-5. Energy band diagrams of Schottky junctions formed between (a) a metal and a p-type semiconductor and (b) a metal and an n-type semiconductor at thermal equilibrium. ......... 44 2-6. A schematic diagram showing the mechanisms by which photons are converted to free carriers in an organic solar cell. (a) Photon absorption in semiconductor layers generates 15 excitons, bound electron-hole pairs. (b) Excitons diffuse around in the material before radiative or non-radiative recombination. (c) When excitons reach the interface between a donor and an acceptor, charge transfer can occur. (d) Charges are spatially separated, transported, and extracted at the electrodes. HOMO and LUMO is the abbreviation of highest occupied molecular orbital and lowest unoccupied molecular orbital, respectively. 46 2-7. Ideal current-voltage characteristics of a solar cell in the dark (black curve) and under AMi.5G illumination (red curve). An ideal solar cell I-V curve in the dark should exhibit strong rectification.............................................................................................................48 2-8. An equivalent circuit diagram of the solar cell comprising a current source (IL), a diode a shunt resistor (Rsh), (Id), and a series resistor (R,). A load resistor (RL, not a part of the solar cell) is shown in the diagram to represent the external load. .................................................. 48 3-1. Schematic of idealized MPTMS monolayer formation on hydroxylated oxide substrate surface with the presence of a physically-absorbed water layer......................................58 3-2. A photograph of the multi-vessel dip coater used to deposit high-quality PbS QD film in this thesis.................................................................................................................................5 9 3-3. Porous and rough film grown by a homogeneous reaction. .............................................. 61 3-4. A photograph of the custom-built Teflon substrate holder. .............................................. 62 3-5. Schematic illustration of overlaid patterned ITO, BCP, and Mg:Ag/Ag electrodes. The overlap between ITO and metal electrodes defines the ten devices with an individual active area of 1.24 mm 2 on a half-inch glass substrate. The same pattern was also employed for PbS QD/CdS bilayer heterojunction devices. The pattern was designed by Dr. Alexi Arango. .......................... ................................................................................................................ 16 65 4-1. A schematic representation of a spray coating setup composed of a handheld airbrush, a gas regulator, and a nitrogen cylinder. Relevant processing parameters are also shown......70 4-2. Optical micrographs of PbS QD film deposited via (a) spin coating, (b) dip coating, and (c) spray coating . .................................................................................................................... 71 4-3. AFM images of PbS QD film deposited via (a) spin coating and (b) dip coating........71 4-4. A three-dimensional profilometry image of spray-coated PbS QD film. ......................... 72 4-5. J-V characteristics of PbS QD Schottky devices under simulated AMI.5G illumination. PbS QD film in these devices was deposited via (a) spin coating, (b) dip coating, and (c) spray co ating . ............................................................................................................................. 72 5-1. (a) Schematic of the PbS QD/CdS heterojunction device fabricated in this study. (b) Energy band diagram showing band edges of isolated 1.5eV PbS QDs and CdS along with electrode w ork functions...................................................................................................................76 5-2. Absorbance spectrum of PbS QDs in solution used in this work. .................................... 77 5-3. SEM im age of the CdS thin film on glass. ........................................................................ 78 5-4. Current-voltage characteristics of PbS QD and CdS thin films. The electrical conductivities of PbS QDs and CdS were measured using a planar interdigitated configuration. For PbS QDs, the channel length is 6 ptm, the channel width is 1.95 cm, and the film thickness is 100 nm. For CdS, the channel length is 120 ptm, the channel width is 0.75 cm, and the film thickness is 120 nm. Consequently, the conductivities of PbS QDs and CdS are -2 x 10-8 and ~1 x 10-3 S/cm , respectively........................................................................................... 17 78 5-5. (a) Cross-sectional SEM image of the device. Platinum was deposited on top of the device before focused ion beam milling to protect the underlying layers. (b) XPS depth profile of a PbS QD/CdS device without Al electrodes. .................................................................. 79 5-6. J-V characteristics of nine PbS QD/CdS heterojunction devices fabricated on the same substrate and measured under simulated AMi.5G solar illumination. The scattered J-V characteristics manifest the device irreproducibility......................................................80 5-7. A representative top-view optical micrograph of a finished PbS QD/CdS device. Multiple pinholes can be identified in the image. ......................................................................... 80 5-8. (a) A representative top-view optical micrograph of finished PbS QD Schottky devices. No pinhole is identified in this image. (b) J-V characteristics of nine PbS QD Schottky devices fabricated on the same substrate and measured under simulated AMI.5G solar illumination show good reproducibility. ........................................................................................... 81 5-9. Representative optical micrographs of (a) as-deposited PbS QD film, (b) PbS QD film with a CdS overlayer and with no pinhole, and (c) PbS QD film with a CdS overlayer and with a pinhole. The pervasive submicron dark spots in (b) and (c) are the CdS crystallites which were initially formed in the bulk of the chemical bath solution and subsequently adsorbed and em bedded in CdS film ............................................................................................. 82 5-10. An optical micrograph of an ITO-coated glass substrate cleaned with the procedure described in Chapter 3 still shows dust particles/contaminants (highlighted with the red circles) adsorbed on the surface ...................................................................................... 82 5-11. Optical micrographs of ITO-coated glass substrates (a) before and (b) after MPTMS treatment. The red circles in (b) highlight the surface contaminants. .............................. 18 83 5-12. An optical micrograph of PbS QD film deposited from a solution of poor colloidal stability. Multiple particle defects can be identified in the image................................................. 84 5-13. Photographs of centrifuged PbS QD solution added with (a) excess and (b) reduced amount of ethanol. Complete purification process should be referred to Chapter 3. The supernatant in the solution with excess amount of ethanol is much clearer than that with reduced amount of ethanol, suggesting that most of the QDs was precipitated out during centrifugation in the solution w ith excess amount of ethanol.......................................................................... 84 5-14. Optical micrographs of the PbS QD/CdS devices fabricated (a) without and (b) with additional measures to prevent pinhole formation. The red circles in (a) highlight the particle defects/pinholes. The blurred black spots in both images were the residual of the QD film on the back side of the substrates. During a dip-coating process, both sides of the substrates were coated, and the film on the back side was wiped off before the measurement. Therefore, these blurred spots are not in the device and can be ignored..........................85 5-15. J-V characteristics of nine PbS QD/CdS heterojunction devices fabricated with the newly-implemented experimental methods on the same substrate and measured under simulated A M 1.5G solar illum ination............................................................................ 85 5-16. J-V characteristics of typical PbS QD Schottky (red) and PbS QD/CdS heterojunction (blue) devices measured in the dark and under AM1.5G simulated solar illumination..............86 5-17. J-V characteristics of an inverted CdS/PbS QD heterojunction solar cell measured in the dark and under AMl.5G simulated solar illumination. To verify the origin of the device rectifying behavior, symmetric top and bottom ITO contacts were employed. Due to some spikes of chemical-bath-deposited CdS thin films, a thick (310 nm) PbS QD layer was used 19 in this inverted device to avoid shorting, which may result in a worse photovoltaic perform ance than a normal device. ............................................................................... 87 5-18. (a) J-V characteristics of representative PbS QD/CdS heterojunction solar cells with varying PbS QD layer thicknesses. (b) Performance characteristics of the above devices. The error bars are evaluated from the standard deviation of 18 devices fabricated with the same condition for each thickness of the PbS QD film...................................................89 5-19. Complex refractive indices of PbS QDs and CdS. ........................................................ 90 5-20. AFM surface images of (a) the PbS QD thin film and (b) the CdS thin film...........91 5-21. Diffuse reflectance spectrum of a full PbS QD/CdS device (without Al electrode)......91 5-22. Measured EQE (red curves) and modeled active-layer absorption spectra (black curves) of PbS QD/CdS devices with varying PbS QD thicknesses: (a) 40 nm, (b) 100 nm, (c) 160 nm, (d) 220 nm , and (e) 280 nm . ......................................................................................... 89 5-23. Modeled photon absorption rate (1/sec-cm 3 ) for PbS QD/CdS heterojunction devices with varying PbS QD thicknesses: (a) 40 nm, (b) 100 nm, (c) 160 nm, (d) 220 nm, and (e) 280 nm. The x- axis (distance) is measured from the ITO surface where light is incident. The horizontal stripes seen in the simulation plots result from the use of AM1.5G solar spectrum as a light source for the calculated absorption/generation rate. There are multiple dips in the solar spectrum due to light absorption by 03, 02, H 20, and CO 2 ................ . . . . . . . . .. . . . . . . . . . . 94 5-24. The modeled total generation rates in PbS QD layers of different thicknesses (offset for clarity)...............................................................................................................................9 5 5-25. The integrated generation rate (1/sec-cm 2 ) as a function of the distance from the PbS QD/CdS interface for the above PbS QD thicknesses.....................................................95 20 5-26. Jsc integrated from the modeled (black curve, assuming 100% IQE) and measured (red squares) EQ E spectra.................................................................................................... 96 5-27. Measured (black square) and modeled (red circle for the lower bounds; blue triangle for the upper bounds) device capacitance at zero-bias as a function of PbS QD thicknesses. The lower bounds were calculated, assuming that both the PbS QD and CdS layer are fully-depleted; while the upper bounds were calculated, assuming that only the PbS QD layer is fully-depleted and none of CdS layer is depleted. .............................................. 97 5-28. Modeled short-circuit current density of the PbS QD/CdS devices with various CdS layer thicknesses. The thickness of PbS QD layer for this modeling is fixed at 170 nm, which is effective carrier extraction range estimated in the previous sections of this thesis. ......... 99 5-29. Simulated loptical electric field12 as a function of wavelength and position for various CdS thicknesses: (a) 0 nm, (b) 50 nm, (c) 180 nm, (d) 280 nm, and (e) 400 nm. The thickness of PbS Q D layer is fixed at 170 nm ...................................................................................... 100 6-1. (a) The schematic device architecture and (b) the corresponding flat-band diagram..........107 6-2. TEM images of (a) P3HT nanofibers and (b) CdS QDs. ................................................... 108 6-3. UV-VIS absorbance spectra of (a) amorphous P3HT, P3HT nanofribers, and (b) CdS QDs. .............................................................................................. .......................................... 10 8 6-4. TEM images of P3HT nanofiber/CdS QD film synthesized (a) without self-assembly (random-mix process) and (b) using a solvent-assisted self-assembly process. The inset images show schematic representations of each; the random-mix and self-assembly method is used to control the interface between CdS QDs (yellow spheres) and P3HT nanofibers (purp le lin es). .................................................................................................................. 21 109 6-5. (a) An UV-VIS absorbance spectrum of self-assembled P3HT nanofiber/CdS QD film. (b) Solution PL spectra of the random-mixed and self-assembled P3HT nanofiber/CdS QD sam p les...........................................................................................................................1 10 6-6. XPS spectra of high resolution S2p in (a) random-mixed and (b) self-assembled P3HT nanofiber/CdS QD film . ............................................................................................... I 1 6-7. Time-resolved photoluminescence spectra of the random-mixed and self-assembled P3HT nanofiber/CdS QD film. The film was excited by a pulsed diode laser (PicoQuant) at 414 nm, 2.5 MHz. Emission from the film was collected by an avalanche photodiode (PicoQuant) with a 460nm band-pass filter. Photoluminescence decay of the sample was measured with a time-correlated single photon counting module (PicoQuant)..................111 6-8. Tapping mode AFM images of P3HT nanofiber/CdS QD thin film (a) before and (b) after EDT ligand exchange. (c) TEM image of curved P3HT NWs after the ligand exchange. . 112 6-9. Charge carrier recombination rate constants krec of pristine and EDT-treated devices determined by transient open circuit voltage decay measurements at different illumination lev els............................................................................................................................... 1 13 6-10. (a) J-V characteristics of an electron only P3HT nanofiber/CdS QD device before and after EDT treatment. (b) J-V characteristics of a hole only P3HT nanofiber/CdS QD device before and after EDT treatment. The effect of EDT treatment and P3HT nanofiber curving on the transport properties of P3HT nanofiber/CdS QD film was demonstrated by the DC J-V characteristics of the electron-only and hole-only devices. The electron-only device consisted of Ag:self-assembled monolayer (SAM), P3HT nanofiber/CdS QD, and Mg:Ag layers. The Ag:SAM layer blocks the hole injection [Error! Reference source not found.]. The treated electron-only device shows a strong enhancement of electron transport over the 22 untreated device. The hole only device consisted of ITO, PEDOT:PSS, P3HT nanofiber/CdS QD, and Au layers, where the Au layer prevents electron injection. The treated device showed a decrease in current density. This result suggests that the EDT treatm ent reduces the hole transport.................................................................................114 6-11. (a) Cross-sectional TEM image of P3HT nanofiber/CdS QD hybrid solar cells. (b) Energy dispersive X-ray spectroscopy elemental mapping (Red color: Cd; Green color: In; and Blue: A g) of the im age show n in (a). ........................................................................................ 6-12. Cross-sectional TEM EDS mapping image ITO/PEDOT:PSS/P3HT:CdS/BCP/Mg:Ag hybrid solar cell device. (a)-(d): of 115 the The Mg, Cd, S, and In elem ent m apping im ages................................................................................... 115 6-13. J-V characteristics of random-mixed (blue) and self-assembled (red) P3HT nanofiber/CdS QD so lar cells..................................................................................................................116 6-14. (a) J-V characteristics of pristine (blue) and EDT-treated (green) P3HT nanofiber/CdS QD solar cells. (b) The photovoltaic performance summary (Jsc and Voc) of the EDT-treated and self-assembled devices as a function of the CdS weight concentration.......................117 23 24 Chapter 1 Motivation 1.1 Solar Photovoltaic Background Since the twentieth century, the living standard of most of the world population has increased in a significant and unprecedented way. However, at the same time, human activity has also casted a huge impact on the earth. Atmospheric carbon dioxide concentration and global surface temperature have risen since 1960 [1, 2], and evidence suggests that the global warming is a result of a rise in greenhouse gas emission caused by human activity [3]. Such change could be unsustainable or even catastrophic, if it is continued in the current fashion. Sustainable development, the development that meets the needs for economy and society without compromising the environment, is therefore one of the great challenges of the twenty-first century. Global greenhouse gas emissions by sector from 1990 to 2005, measured in carbon dioxide equivalents, shows that the energy sector accounts for two thirds of total anthropogenic carbon dioxide emissions [4]. As a result, in order to decrease global greenhouse gas emissions, people cannot ignore the major contribution from the energy sector, including both energy production and consumption. In addition, although hydraulic-fracturing technologies have recently opened up considerable oil and gas resources, these reserves could last for only about sixty years (they may be extended by new technologies and unconventional deposits) at current consumption rates. Coal reserves will last for at least a century at current rates of consumption [5]. Nevertheless, none of these energy resources are renewable and will be exhausted in the foreseeable future. To 25 summary, energy plays an important role for sustainable development. We need not only reliable and affordable energy sources to continue economical and social development but also clean energy sources to mitigate the adverse effects produced during energy production and consumption on the environment. Among various energy sources, solar energy is particularly abundant [6], and solar photovoltaics is relatively clean compared to coal-fired and natural gas power plants regarding greenhouse gas emission [7]. For the reason given above, solar photovoltaics has been growing rapidly in this past decade [8]. The global cumulative installed solar photovoltaic capacity has surpassed 100 GW in 2012, achieving just over 101 GW [9]. However, photovoltaics still enjoys marginal portions of the overall power generation. In 2010, solar photovoltaics accounted for only 0.06% of total energy consumption [8]. To accelerate the deployment of solar photovoltaics, their cost must be reduced such that solar electricity can be as cheap as the electricity generated from conventional sources such as fossil and nuclear fuels. Table 1-1 lists solar electricity price in US cents per kilowatt-hour reported in 2012 for various installation scales and climate conditions. Even the lowest price (15.15 cents/kWh) of solar electricity generated in the industrial scale and sunny climate condition is still well above that (~10 cents/kWh) of conventional electricity in most of the states in the United States in 2012 [11]. The cost of solar electricity can be divided into a number of components, including photovoltaic modules, electrical and structural balance of systems, inverters, labors, and other soft costs. As of 2013, photovoltaic modules still account for the largest portion (32%) of the solar electricity cost [12]. It is thus critical to cost down photovoltaic modules to decrease overall solar electricity cost. 26 ResdenialSunny Residential Table 1-1. 28.91 Installed System Cloudy 63.60 Commercial Sunny Installed System 19.42 Cloudy 42.73 Industrial Sunny 15.15 Installed System Cloudy 33.32 Price of solar electricity in US cents per kilowatt-hour reported in 2012 for various installation scales and climate conditions [10]. The residential price is calculated based upon a standard 2 kilowatt peak system; the commercial price, a 50 kilowatt peak system; the industrial price, 500 kilowatt peak system. 1.2 Low-Temperature Solution-Processed Colloidal Quantum Dot Solar Cells Low-temperature solution-processed solar cells have the potential to decrease manufacturing cost for photovoltaic modules. These solar cells can be built on lightweight and flexible substrates and are compatible with roll-to-roll manufacturing process such as printing. In addition, there is no vacuum process required to manufacture these solar cells. These characteristics give rise to not only low manufacturing cost but also high manufacturing throughput and short energy-payback time. Because these solar cells can be built on lightweight 27 and flexible substrates, the transportation and installation cost for these photovoltaic panels will be low as well. Organic semiconductors and colloidal quantum dots are the two main material systems that can be incorporated into low-temperature solution-processed solar cells. Colloidal quantum dots are particularly interesting for photovoltaic applications because of their spectral tunability and will be the focus of this thesis. Colloidal quantum dots (QDs) are chemically synthesized semiconductor nanocrystals coated with organic ligands (Figure 1-1) which passivate the surface of nanocrystals and provide solubility to nanocrystals in various organic solvents. The particle size of QDs can be well controlled via synthetic conditions which will be detailed in Chapter 2. When the particle size is smaller than the exciton Bohr radius [13] of the constituent semiconductor, the allowed energy levels for excitons (or electrons and holes) become quantized such that the band gap energy of the semiconductor changes with the particle size, which is termed "quantum confinement effect." The relationship between the band gap energy and particle size can be simplified and understood with a particle in a spherical box quantum mechanical model (Eg lI/d 2 , Eg = band gap energy, and d = particle diameter) as illustrated in Figure 1-2. As a consequence, the band gap energy of QDs can be tuned by controlling their particle size. Inorganic semiconductor core Organic ligand shell Figure 1-1. Schematic of a colloidal quantum dot with inorganic semiconductor core and organic ligand shell. 28 Particle in a box Conduction Band Valence Band Decreasing particle sizes Increasing band gaps Figure 1-2. The effect of quantum confinement on the energy levels of electrons (solid circle) and holes (open circle) in semiconductor nanoparticles. Table 1-2 summarizes theoretical power conversion efficiency (Shockley-Queisser limit [14, 15]) of single- and multijunction solar cells and the corresponding optimal band gap energy for the constituent junctions. Multijunction solar cells, consisting of multiple semiconductors of different band gap energy, extract solar energy more efficiently and promise for higher efficiency. For example, power conversion efficiency increases from 31% to 49% when triple junctions, instead of single junction, are employed. However, since multiple semiconductors are involved in the multijunction solar cells, considerable effort has to be exerted to optimize the properties of constituent materials, which significantly increases the manufacturing complexity and cost of these cells. Band gap tunability of QDs offers the opportunity to realize multijunction solar cells 29 with single semiconductor material, providing the potential to offset the manufacturing disadvantages derived from multijunction devices. As shown in Figure 1-3, PbS QDs covers all the optimal band gap energies for triple-junction solar cells. Number of junctions 23 Band gap 1100 nm 1320 nm 760nm 760 nm 1750 170nnm 1070 nm 680 nm Efficiency 31% 44% 49% Table 1-2. Theoretical power conversion efficiency of single- and multijunction solar cells and the corresponding optimal band gap energies for the constituent junctions. Band gap energies are expressed with equivalent photon wavelength. 6 5 7 UJ C 0 4 i2 0400 800 1200 1600 2000 Wavelength (nm) Figure 1-3. Absorption spectra of PbS QDs with a series of particle diameters. The first excitonic absorption peak of PbS QDs can be tuned readily from 600 to 2200 nm [16-18]. On the other hand, the band gap of bulk PbS is 0.41 eV, which is approximately 2958 mu. It has to be noted that although the band gap of Cu(In,Ga)Se 2 , an extensive studied semiconductor for a thin film solar cell, can be adjusted by substitution of Ga for In, its band gap changes from 1.04 to 1.67 eV (or 1195 to 744 nm) [19] and does not cover most of the band gaps 30 calculated for multijunction solar cells (Table 1-2). 1.3 Thesis Outline This thesis focuses on the electron-hole separation junction (or interface), one of the key elements in solar cells. Specifically, it tackles inorganic-inorganic bilayer heterojunctions and inorganic-organic bulk heterojunctions; both of which involve QDs. Chapter 2 first gives an overview of QD synthesis then reviews electrical properties (carrier mobility, carrier type, and carrier concentration) of QDs and solar cell device physics, including photovoltaic mechanisms in both inorganic and organic solar cells, and current-voltage characteristics of solar cells. The chapter provides the chemistry and physics background for colloidal QD-based solar cells. Chapter 3 elaborates the experimental procedures for the material synthesis and the fabrication and characterization of PbS QD/CdS bilayer heterojunction devices and poly(3-hexylthiophene-2,5-diyl) nanofiber/CdS QD bulk heterojunction devices. Chapter 4 presents PbS QD Schottky devices fabricated by spin coating, dip coating, and spray coating. Comparison of film morphology and device performance in the spin-, dip, and spray-coated devices is reported. This chapter paves the way for the fabrication of PbS QD photovoltaic devices with high yield and satisfying reproducibility. Chapter 5 describes the realization, optimization and characterization of PbS QD/CdS bilayer heterojunction devices, the first low-temperature solution-processed pn heterojunction devices employing PbS QDs. Through a combination of characterization techniques, fundamental material and device parameters are reported. Chapter 6 presents a facile method to improve the power conversion efficiency of 31 poly(3-hexylthiophene-2,5-diyl) nanofiber/CdS QD bulk heterojunction solar cells through the modification of the interfacial interaction between organic electron donors and inorganic electron acceptors. Chapter 7 summarizes the findings made from this thesis work and briefs the future work that can be done to further improve the efficiency of QD-based photovoltaic devices. 1.4 Reference Source: Met Office Hadley Centre; NOAA; Scripps Institute of Oceanography; Sydney 1 Levitus et al, GRL 2 Climate change: The measure of global warming, The Economist, 2013 3 Etheridge, D. M.; Steele, L. P.; Langenfelds, R. L.; Francey, R. J.; Barnola, J.-M. & Morgan, V. I., Natural and anthropogenic changes in atmospheric C02 over the last 1000 years from air in Antarctic ice and firn, Journal of Geophysical Research: Atmospheres, 1996, 101, 4115-4128 4 Climate Analysis Indicators Tool, World Resources Institute, 2009 5 Jeff Tollefson & Richard Monastersky, Nature, 2012, 491, 654-655 6 Graetzel, M.; Janssen, R. A. J.; Mitzi, D. B. & Sargent, E. H., Materials interface engineering for solution-processed photovoltaics, Nature, 2012, 488, 304-312 7 U.S. Department of Energy, Report on the First Quadrennial Technology Review, 2011 8 REN2 1, Renewables Global Status Report, 2012 9 European Photovoltaic Industry Association, 2013. 10 Solarbuzz Retail Pricing Environment, 2012 11 U.S. Energy Information Administration, 2012 32 12 GTM Research, Solvida Energy Group, Inc., 2012 13 Excitons are bound electron-hole pairs. The exciton Bohr radius is the avarage distance between the electron and hole of an exciton in its ground state. 14 Shockley, W. & Queisser, H. J.Detailed Balance Limit of Efficiency of p-n Junction Solar Cells, Journal of Applied Physics, 1961, 32, 510-519 15 Sargent, E. H., Infrared photovoltaics made by solution processing, Nat Photon, 2009, 3, 325-33 16 Hines, M. & Scholes, G., Colloidal PbS Nanocrystals with Size-Tunable Near-Infrared Emission: Observation of Post-Synthesis Self-Narrowing of the Particle Size Distribution, Adv. Mater., 2003, 15, 1844-1849 17 Liu, T.-Y.; Li, M.; Ouyang, J.; Zaman, M. B.; Wang, R.; Wu, X.; Yeh, C.-S.; Lin, Q.; Yang, B. & Yu, K., Non-Injection and Low-Temperature Approach to Colloidal Photoluminescent PbS Nanocrystals with Narrow Bandwidth, The Journal of Physical Chemistry C, 2009, 113, 2301-2308 18 Moreels, I.; Justo, Y.; De Geyter, B.; Haustraete, K.; Martins, J. C. & Hens, Z., Size-Tunable, Bright, and Stable PbS Quantum Dots: A Surface Chemistry Study, ACS Nano, 2011, 5, 2004-2012 19 Wei, S.-H.; Zhang, S. B. & Zunger, A., Effects of Ga addition to CuInSe 2 on its electronic, structural, and defect properties, Applied Physics Letters, 1998, 72, 3199-3201 33 Chapter 2 Basics of Colloidal Quantum Dot Solar Cells This chapter is to provide introductory and theoretical background for colloidal QD solar cells, including colloidal QD synthesis, optical and electrical properties of QDs and methods to tune these properties, and solar cell device physics and electrical characteristics. 2.1 Colloidal Quantum Dot Synthesis The investigation of colloidal quantum dots began in early 1980s, when CdS nanocrystals with broad size distribution were precipitated and grown via Ostwald ripening in aqueous solution [1]. A breakthrough in synthesizing monodisperse colloidal QDs was made by Chris Murray, David Norris, and Moungi Bawendi in 1993, when hot injection method [2] was first introduced. Since then, there have been numerous studies attempting to improve size distribution and optical property of colloidal QDs and also change the morphology, composition, and chemistry of colloidal QDs. However, hot injection method remains the mainstream colloidal QD synthesis protocol because of the narrow size distribution it can provide. Therefore, colloidal QDs in this thesis are synthesized via hot injection method. Detailed description of synthesis procedure for colloidal QDs of different compositions will be provided later in the corresponding chapters. Here, a generalized colloidal QD synthesis is presented. Colloidal QD synthesis is performed with a Schlenk line which provides air- and moisture-free environment. Figure 2-1 shows the photograph of the setup used for colloidal QD synthesis. One precursor compound is first dissolved in a heated high-boiling point solvent contained and degassed in a three-neck flask. The precursor solution is further heated to the reaction temperature. Then, another degassed precursor solution is swiftly injected to the flask, 34 which raises the precursor concentration above the nucleation threshold and leads to a nucleation burst. Figure 2-2 illustrates precursor concentration as the reaction evolves according to LaMer colloid growth model [3]. Rapid injection ensures that most of nuclei formed at the same time. In an ideal case, the lowering of precursor concentration due to the nucleation event and the reducing of overall solution temperature due to the addition of a room-temperature precursor solution terminates nucleus formation. Precursors are then consumed by the nanocrystal growth. The quenching of nucleus formation prior to or at the early stage of nanocrystal growth is a key for narrow particle size distribution achieved with hot injection colloidal QD synthesis because it allows all nanocrystals to grow at the same time. Eventually, the reaction will reach equilibrium between material in nanocrystals and precursor monomers, and the size distribution will start to increase because big particles grow at the expense of small particles' shrinkage known as Ostwald ripening. Particle size can be controlled via various synthesis conditions, such as precursor reactivity, coordinating/non-coordinating solvent ratio, growth temperature, and growth period. When more reactive precursor is used, more nuclei are formed, resulting in smaller particle size and larger size distribution [4]. Coordinating solvent promotes precursor consumption by nanocrystal growth, which leads to larger particle size [5]. Higher growth temperature and longer growth period also increase particle size. 35 Figure 2-1. A photograph of colloidal QD synthesis setup, including a Schlenk line, a three-neck flask, a condenser, a stirrer, a temperature controlling system, an injection gun, and a mechanical rotary pump (not shown in the picture). The black injection gun shown in this picture is used for large-scale (gram-scale) synthesis, while for a conventional small-scale synthesis (100 milligram-scale) synthesis, a syringe is used. C 0 C U o Z .0 C 4. Nucleation threshold = U ~ (U + Ostwald ripening Time Figure 2-2. Cartoon depicting the precursor concentration evolution along different stages of colloidal QD nucleation and growth under the framework of LaMer colloid growth model. 36 2.2 Electrical Properties of Colloidal Quantum Dots After synthesis, colloidal QDs are capped with long-chain organic ligands, such as oleic acid, oleylamine, trioctylphosphine, and trioctylphosphine oxide, which provide solubility to QDs in a range of organic solvents. The capping ligands also play an important role in the electrical properties of colloidal QDs, such as carrier mobility, carrier type and concentration, and carrier traps. 2.2.1 CarrierMobility In a thin film composed of colloidal QDs, carriers are transported via either tunneling or hopping mechanisms which are largely affected by the distances between adjacent QDs. Therefore, a lot of effort has been made to shrink the inter-QD distances to improve carrier mobility. This is generally achieved via solution- or solid-phase chemical treatments by which long-chain organic ligands are replaced by short organic or inorganic ligands. Although carrier mobility can also be increased by a high-temperature annealing process which sinters QD thin film [6], the quantum confinement effect is lost and thus is not within the focus of this thesis. Solution-phase chemical treatment here is a pre-deposition ligand exchange process occuring in a QD solution and requires the preservation of colloidal stability of QDs after ligand exchange for the following thin film deposition with small roughness. Common organic molecules for solution-phase ligand exchange are pyridine [7] and n-butylamine [8]. Recently, Dmitri Talapin's group developed molecular metal chalcogenide complex ligands and achieved electron mobility as high as 16 cm 2/Vs for CdSe colloidal QDs with In 2 Se2 ligands [9]. Cherie Kagan's group also demonstrated thiocyanate-capped CdSe QD thin film with high electron mobility of 1.5 ± 0.7 cm 2/Vs [10]. Both of these values are acquired by field effect transistor 37 measurement and are considered very high for non-sintered colloidal QD thin film. Although the physical size of the ligands used in these two cases are tiny, these ligands charge QD surfaces and provide electrostatic repulsion between QDs, which prevents aggregation and offers solubility in polar solvents. Solid-phase chemical treatment, on the contrary, is a post-deposition ligand exchange process carried out after QD thin film is deposited. Small ligands, such as 1,2-ethanedithiol [11], 1,4-benzenedithiol [12], 3-mercaptopropionic acid [13], and hydrazine [14], used in this process drastically shorten inter-QD distances and insolubilize QDs. Untreated oleic acid capped-PbS QDs with hole field-effect mobility of 3 x 10- cm 2/Vs is reported [15]. After treated with 1,2-ethanedithiol, 1,4-benzenedithiol, and 3-mercaptopropionic acid, hole field-effect mobility is increased to 1.7 x 10-3 3.8 x 10-, and 2.7 x 1 0 -3 cm 2 /Vs, respectively [16]. In addition, Yao Liu et al. have also shown that electron and hole field-effect mobility decrease exponentially with increasing ligand length [17]. Recently, Edward Sargent's group has also demonstrated that with halide treatments, such as cetyltrimethylammonium bromide, the electron field-effect mobility of PbS QDs can be improved to 4.4 x 102 cm 2/Vs [18]. Because of the large volume shrinkage, cracks usually form in the treated QD film as shown in Figure 2-3. Therefore, layer-by-layer coating techniques are employed to fill the cracks generated in the underlying layers to prevent shorts. 38 Figure 2-3. A Scanning electron microscopy image of a treated QD film. Cracks are usually developed in thick QD film after post-deposition treatment. In summary, carrier mobility in QD film is strongly correlated with the capping ligands: the shorter the ligand, the higher the mobility. Various solution- and solid-phase chemical treatments are implemented to replace bulky native ligands with short ligands, providing a pathway toward solution-processed high-performance electronic and optoelectronic devices. 2.2.2 Carrier Type and Density Carrier type and density are important factors for the operation of solar cells as will be described later in this chatper. It has been reported in several QD material systems that carrier type and density can be controlled via chemical treatment and are also sensitive to the environment to which QDs exposed. Conduction of PbSe QDs can be switched from p- to n-type by hydrazine post-deposition treatment and reversed to p-type as the hydrazine desorbed [14]. By increasing the doping level, silver can be changed from n- to p-type dopants for CdSe QDs [19]. N-type doping of CdSe QDs can also be realized by thermal diffusion of indium [9], and electron transfer from sodium 39 biphenyl [20]. Metal chalcogenide complexes allow switching of CdTe QD conduction from p-type to ambipolar to n-type [21]. Pre-deposition incorporation of cadmium switches conduction of InAs QDs from n- to p-type [22]. Chemical treatments, including organic and halide ligands, are employed to tailor doping type and concentration of PbS QDs [23, 24, 25]. Oxidation is found to increase p-type doping of PbS QDs [15]. In addition, carrier type and density in PbSe QDs are shown to be connected to QD's stoichiometry: Pb-rich QDs are n-type; Se-rich, p-type [26]. 2.3 Solar Cell Device Physics A solar cell is a semiconductor diode which directly converts sun light into electricity. In a conventional inorganic solar cell, this conversion process primarily consists of photon absorption (carrier generation) and carrier separation. While in an excitonic solar cell, such as an organic solar cell [27], very similar but not exactly identical processes are involved, namely photon absorption (exciton generation), exciton diffusion, charge transfer, charge separation, carrier transport and carrier extraction [28]. Since this thesis work involves both inorganic and organic solar cells, in the following sections I will first describe the details of these processes in conventional inorganic solar cells, followed by the explanation of similar conversion process in organic solar cells. 2.3.1 Photovoltaic Processes in Inorganic Solar Cells Photon Absorption and CarrierGeneration: When light shines on a solar cell, photons, if not reflected from the solar cell, with energy equal to or greater than the band gap of semiconductors can be absorbed. Such photon absorption 40 event excites electrons from valence bands to conduction bands and generates free carriers (electrons and holes). However, not all the photons with energy equal to or greater than the band gap of semiconductors are absorbed. In materials with low absorption coefficient or small thickness, photons can transmit without being absorbed, even though their energy is sufficient to excite electron from valence bands to conduction bands. Therefore, reflection and transmission are two of the loss mechanisms that need to be eliminated or minimized to achieve high photocurrent density and high device performance. The number of free carriers generated by photon absorption is limited by the flux of photons with enough energy to excite electrons from valence bands to conduction bands. The smaller the band gap of semiconductors, the more free carriers can be generated by photon absorption. However, because carriers generated by high-energy photons (with energy greater than the band gap of semiconductors) relax rapidly to the band edge of semiconductors and also because of energy conservation, the usable energy is smaller when semiconductors of smaller band gaps are employed. Due to this trade-off, to realize high device efficiency, there is an optimal band gap for the semiconductor used in a solar cell as mentioned in chapter one. Although there are approaches proposed to exploit or better use photons with energy smaller or greater than the band gap of semiconductors, such as an intermediate-band solar cell [29], a multiple exciton generation solar cell [30], and a hot carrier solar cell [31], these types of device are still in their very early stage of research and rather immature and thus are beyond the scope of this thesis. Carrier Separation: After photogeneration, carriers need to diffuse and drift to the corresponding electrodes to perform electrical work to the loads in an external circuit. The relevant parameters for diffusion 41 and drift processes are minority carrier diffusion length and depletion width. For the carriers generated outside of the depletion region, they need to diffuse to the depletion region, but the distance they can diffuse before annihilated via recombination is limited by the minority carrier diffusion length. For the carriers generated in or successfully diffusing into the depletion region, they can be swept by the built-in electric field to the corresponding electrodes and contribute to electric current. Built-in electric field is typically formed in a semiconductor-semiconductor pn junction or a metal-semiconductor Schottky junction which forms the basis of a solar cell and is elaborated in the following paragraphs. When p-type and n-type semiconductors are joined together, due to the electron and hole concentration gradient across the p-type and n-type semiconductor junction, holes in the p-type semiconductor diffuse to the n-type semiconductor, and electrons in the n-type semiconductor diffuse to the p-type semiconductor to establish chemical potential equilibrium. The diffusion of free carriers leaves behind uncompensated positively charged donor atoms in the n-type semiconductor and negatively charged acceptor atoms in the p-type semiconductor. The charged atoms created the so-called built-in electric field (or built-in potential, Vbi, or energy band bending) which sweeps away free carriers in the junction region, and the junction region where free carriers are depleted is named depletion region whose width is denoted as Wd and equal to: q2 W4/c ==[2-cs-o where es, co, NA+ND (NA+ND) Vbi bi]1/2 21 21 q, NA, and ND are the dielectric constant of the semiconductor, vacuum permittivity, elementary charge, acceptor concentration and donor concentration, respectively. Figure 2-4 shows the energy band diagram of a pn homojunction at thermal equilibrium. The depletion width, WP, in the p-type semiconductor and the depletion width, W,, in the n-type semiconductor are given by: 42 WP = [Iq Wn = ND NA(NA +ND) Vbi ND (NA+ND) bi 1/2 1/2 (2-2) (2-3) It can be seen from equation 2-2 and 2-3 that the depletion region is not evenly distributed in the p-type and n-type semiconductors; instead, it spreads mostly in the less-doped side of the semiconductor (either in the p-type or n-type semiconductor). As a result, in heterojunction solar cells such as CdTe/CdS and CIGS/CdS solar cells [32, 33], it is preferred to have the absorber layer less doped than the window or buffer layer to extend most of the depletion width in the absorber layer and facilitate carrier separation. depletion region EC electrons EC EF holes EF w W p-type n-type Figure 2-4. Energy band diagram of a pn homojunction at thermal equilibrium. When a metal and semiconductor are brought into contact, energy band bending and depletion regions can also be established as in the case of pn semiconductor junctions. Such metal-semiconductor junction is known as a Schottky junction in which the metal can be conceived as a very heavily doped semiconductor and essentially all the depletion region resides in the semiconductor side as depicted in Figure 2-5. 43 (a) EF---- - - - - - --- - EF W metal p-type (b) EC EF - - - - - - - - - - - - - - - - -EF Ev W n-type metal Figure 2-5. Energy band diagrams of Schottky junctions formed between (a) a metal and a p-type semiconductor and (b) a metal and an n-type semiconductor at thermal equilibrium. 2.3.2 Photovoltaic Processes in Organic Solar Cells Conversion of light to electricity in an organic solar cell takes on the mechanistic steps very similar to the ones in an inorganic solar cell but with some differences due to the material properties of organic semiconductors. Exciting organic semiconductors with light does not generate appreciable free electron-hole 44 pairs immediately. In conjugated polymers, only 10% of photoexcitations with energy above the band gap directly create free charges [34]. The majority of the photoexcitations in organic semiconductors forms coulombically bound electron-hole pairs, called excitons, due to small dielectric constants and resultant large exciton binding energy (estimated to be 0.4-0.5 eV [35]) in these materials [27]. These excitons have to diffuse to dissociation sites to be broken apart into free carriers. The dissociation process can be achieved by blending two organic semiconductors of different energy levels together such that it is energetically favorable to transfer electrons from a donor material to an acceptor material and holes from an acceptor material to a donor material. Although the transferred electrons and holes (charge-transfer exitons) can still attract each other, these excitons have decreased binding energy because of increased separation between electrons and holes and thus more susceptible to dissociation under electric fields or thermal activation. The spatially separated charges then need to transport to and be extracted at an anode and cathode in order to contribute to photocurrents. The sequential photovoltaic process starting from photon absorption to carrier extraction in an organic solar cell is schematically illustrated in Figure 2-6. 45 (b) (a) Energy Donor LUMO Acceptor I I I LUMO | I I I LUMO HOMO HOMO HOMO (d) (c) LUMO Donor LUMO Acceptor LM LUMO HOMO HOMO__ HOMO IHOMO Figure 2-6. A schematic diagram showing the mechanisms by which photons are converted to free carriers in an organic solar cell. (a) Photon absorption in semiconductor layers generates excitons, bound electron-hole pairs. (b) Excitons diffuse around in the material before radiative or non-radiative recombination. (c) When excitons reach the interface between a donor and an acceptor, charge transfer can occur. (d) Charges are spatially separated, transported, and extracted at the electrodes. HOMO and LUMO is the abbreviation of highest occupied molecular orbital and lowest unoccupied molecular orbital, respectively. 2.3.3 Current-Voltage Characteristics of Solar Cells Although the photovoltaic processes involved in inorganic and organic solar cells are different, current-voltage characteristics of these devices are identical. Figure 2-7 presents ideal 46 current-voltage (I-V) curves of a solar cell in the dark and under AM1.5G illumination. When a forward bias is applied on a solar cell in the dark, carriers are injected across the potential barriers at a semiconductor-semiconductor or a metal-semiconductor interface such that the current flowing through the device increases exponentially with a positive bias as described in the Shockley diode equation: Id= Io SqV (en' - 1) (2-4) where Id is the current flowing through the diode, 1o is the reverse saturation current, n is the ideality factor, k is the Boltzmann constant, and T is the absolute temperature. Under illumination, a photocurrent (IL) is generated and can be superimposed to the diode current, shifting the diode current downward in a current-voltage plot: I = JO (en qV - 1) - 1, (2-5) Current, voltage, and thus output power is available from the device with the power conversion efficiency (PCE) defined as the ratio of maximum output power to input power and is a critical figure of merit for a solar cell. Input power is the product of device area and the incident light intensity, and the standard incident solar spectrum at the Earth's surface is AM1.5G with integrated power density equal to 100 mW/cm 2. Maximum output power (Po 0 t) is the product of optimal current (Im.) and voltage (Vm,) and can be expressed as: Pout = Isc X Voc X FF (2-6) where ISc is the short-circuit current, VOC is the open-circuit voltage, and FF is the fill factor. Fill factor is the ratio of the optimal current and voltage product to the short-circuit current and open-circuit voltage product, defining the "squareness" of the current-voltage curve in the fourth quadrant of a solar cell under illumination. 47 -- AM1.5G Dark 41-J Voltage Figure 2-7. Ideal current-voltage characteristics of a solar cell in the dark (black curve) and under AM1.5G illumination (red curve). An ideal solar cell I-V curve in the dark should exhibit strong rectification. However, a solar cell in reality is not merely composed of an ideal diode, and thus the I-V curve can deviate appreciably from the one depicted in the Figure 2-6. To account for the non-ideality, shunt resistance (Rsh) and series resistance (R,) are included in the equivalent circuit diagram (Figure 2-8) used to describe the electrical behavior of the device. The components in the equivalent circuit diagram are expatiated in the following paragraph. RS ld IshI L sh V RL -0- Figure 2-8. An equivalent circuit diagram of the solar cell comprising a current source (IL), a diode (Id), a shunt resistor (Rh), and a series resistor (R,). A load resistor (RL, not a part of the solar cell) is shown in the diagram to represent the external load. 48 The current source generates current (IL) from the photogenerated carriers before any recombination loss. The diode, which is the basis of a solar cell as mentioned in the previous section, gives rise to the rectifying current-voltage characteristic. To maximize the current flowing to the external load (RL), shunt resistance must be maximized, whereas series resistance has to be minimized. Thin film defects, such as pinholes and cracks, and recombination in the device result in small shunt resistance, while series resistance is determined by device electrical conductance that is affected by carrier mobility and film thickness. Because the diode is not conductive at zero or small bias (V), and because shunt resistance is much greater than series resistance, shunt resistance can be approximated by the inverse slope of the I-V curve at zero bias. On the other hand, because the diode becomes very conductive at a bias greater than the turn-on voltage of the diode, series resistance can be approximated by the inverse slope of the I-V curve at this bias. Since the slopes of the I-V curve at zero and a turn-on bias determine the "squareness" of the curve, Rsh and R, influence the fill factor of a solar cell. From the viewpoint of energy conservation, the upper limit of open-circuit voltage is the band gap or the effective band gap (conduction band-valence band offset of n- and p-type semiconductors or LUMO-HOMO offset of acceptors and donors) of the semiconductors employed in a homojunction or heterojunction solar cell, respectively. However, in reality V,, in a homojunction solar cell is limited by the built-in potential because it is the driving force for charge separation. From the equation derived below, it can be shown that V. in either a homojunction or a heterojunction solar cell is further reduced by recombination and shorts. Using Kirchhoff's circuit laws, the current (I) and voltage (V) in Figure 2-8 can be related as: (IL - Ia - I)RSh = V + IRS (2-7) Replacing Id in equation (2-7) with the Shockley diode equation (2-4) in which V = V - IRs and 49 rearranging the equation, the current (I) can be written as: V L 1 = Rsh V-IRs ( - enkT / -) (2-8) Rsh With the assumption that R, << Rsh in a solar cell and I = 0 at open circuit condition, the dependence of open-circuit voltage on the reverse saturation current (Io) and shunt resistance (Rsh) can be obtained: V c= In(L-Voc/Rsh Therefore, higher recombination, meaning larger lo and smaller Rsh, decreases the VOc. With more shorts, meaning smaller Rsh, the V., will be lower as well. The effect of recombination and shorts on V.c can also be understood by considering the competition between dark current (Id + Ish) and photocurrent (IL). Since at open circuit condition, the total current (I) equals to zero, dark current can be regarded as a counterbalance to photocurrent: the higher the dark current, the lower the Vc. Consequently, recombination and shorts, which can increase dark current, will reduce Vc. Short-circuit current (Isc), the current (I) at short circuit condition (V = 0), is the photocurrent (IL) subtracted by the diode current (Id) and the shunt current (Ish) as can be seen from Figure 2-8. As mentioned before, the diode current is negligible at zero bias, so it is favorable to keep the shunt current small in order to extract photocurrent as much as possible. Photocurrent is determined by the number of photogenerated carriers or split excitons as discussed in previous section 2.2.1 and 2.2.2 and thus is also proportional to light intensity. 2.4 Conclusion Colloidal QDs are the building blocks of the photovoltaic devices studied in this thesis. The 50 property of these QDs can significantly influence solar cell performance and hence needs to be controlled during or after synthesis. This chapter begins by reviewing the synthesis of colloidal QDs with the emphasis on the control of particle size, the parameter directly determining the band gap of QDs. Then, strong dependence of QD electrical properties, namely carrier mobility, carrier type, and carrier density, on postsynthesis chemical treatments is enumerated to manifest the diversity of colloidal QDs. Finally, relevant device physics and electrical characteristics of inorganic and organic solar cells are introduced in preparation for better understanding the behaviors of the QD-based photovoltaic devices fabricated in the following studies. 2.5 Reference I Rossetti, R.; Nakahara, S. & Brus, L. E., Quantum size effects in the redox potentials, resonance Raman spectra, and electronic spectra of CdS crystallites in aqueous solution, The Journal of Chemical Physics, 1983, 79, 1086-1088 2 Murray, C. B.; Norris, D. J. & Bawendi, M. G., Synthesis and characterization of nearly monodisperse CdE (E = sulfur, selenium, tellurium) semiconductor nanocrystallites, Journal of the American Chemical Society, 1993, 115, 8706-8715 3 LaMer, V. K. & Dinegar, R. H. Theory, Production and Mechanism of Formation of Monodispersed Hydrosols, Journal of the American Chemical Society, 1950, 72, 4847-4854 4 Allen, P.; Walker, B. & Bawendi, M., Mechanistic Insights into the Formation of InP Quantum Dots, Angewandte Chemie International Edition, 2010, 49, 760-762 5 Abe, S.; Capek, R. K.; De Geyter, B. & Hens, Z., Reaction Chemistry/Nanocrystal Property Relations in the Hot Injection Synthesis, the Role of the Solute Solubility, ACS Nano, 2013, 7, 943-949 51 6 Jasieniak, J.; MacDonald, B. I.; Watkins, S. E. & Mulvaney, P., Solution-Processed Sintered Nanocrystal Solar Cells via Layer-by-Layer Assembly, Nano Letters, 2011, 11, 2856-2864 7 Lokteva, I.; Radychev, N.; Witt, F.; Borchert, H.; Parisi, J. & Kolny-Olesiak, J., Surface Treatment of CdSe Nanoparticles for Application in Hybrid Solar Cells: The Effect of Multiple Ligand Exchange with Pyridine, The Journal of Physical Chemistry C, 2010, 114, 12784-12791 8 Johnston, K. W.; Pattantyus-Abraham, A. G.; Clifford, J. P.; Myrskog, S. H.; Hoogland, S.; Shukla, H.; Klein, E. J. D.; Levina, L. & Sargent, E. H., Efficient Schottky-quantum-dot photovoltaics: The roles of depletion, drift, and diffusion, Applied Physics Letters, 2008, 92, 122111 9 Lee, J.-S.; Kovalenko, M. V.; Huang, J.; Chung, D. S. & Talapin, D. V., Band-like transport, high electron mobility and high photoconductivity in all-inorganic nanocrystal arrays, Nat Nano, 2011, 6, 348-352 10 Fafarman, A. T.; Koh, W.-k.; Diroll, B. T.; Kim, D. K.; Ko, D.-K.; Oh, S. J.; Ye, X.; Doan-Nguyen, V.; Crump, M. R.; Reifsnyder, D. C.; Murray, C. B. & Kagan, C. R., Thiocyanate-Capped Nanocrystal Colloids: Vibrational Reporter of Surface Chemistry and Solution-Based Route to Enhanced Coupling in Nanocrystal Solids, Journal of the American Chemical Society, 2011, 133, 15753-1576 11 Luther, J. M.; Law, M.; Song, Q.; Perkins, C. L.; Beard, M. C. & Nozik, A. J., Structural, Optical, and Electrical Properties of Self-Assembled Films of PbSe Nanocrystals Treated with 1,2-Ethanedithiol, ACS Nano, 2008, 2, 271-280 12 Koleilat, G. I.; Levina, L.; Shukla, H.; Myrskog, S. H.; Hinds, S.; Pattantyus-Abraham, A. G. & Sargent, E. H., Efficient, Stable Infrared Photovoltaics Based on Solution-Cast Colloidal Quantum Dots, ACS Nano, 2008, 2, 833-840 52 13 Pattantyus-Abraham, A. G.; Kramer, I. J.; Barkhouse, A. R.; Wang, X.; Konstantatos, G.; Debnath, R.; Levina, L.; Raabe, I.; Nazeeruddin, M. K.; Gratzel, M. & Sargent, E. H., Depleted-Heterojunction Colloidal Quantum Dot Solar Cells, ACS Nano, 2010, 4, 3374-3380 14 Talapin, D. V. & Murray, C. B., PbSe Nanocrystal Solids for n- and p-Channel Thin Film Field-Effect Transistors, Science, 2005, 310, 86-89 15 Klem, E. J. D.; Shukla, H.; Hinds, S.; MacNeil, D. D.; Levina, L. & Sargent, E. H., Impact of dithiol treatment and air annealing on the conductivity, mobility, and hole density in PbS colloidal quantum dot solids, Applied Physics Letters, 2008, 92, 212105 16 Bisri, S. Z.; Piliego, C.; Yarema, M.; Heiss, W. & Loi, M. A., Low Driving Voltage and High Mobility Ambipolar Field-Effect Transistors with PbS Colloidal Nanocrystals, Advanced Materials, 2013 17 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 Letters, 2010, 10, 1960-1969 18 Tang, J.; Kemp, K. W.; Hoogland, S.; Jeong, K. S.; Liu, H.; Levina, L.; Furukawa, M.; Wang, X.; Debnath, R.; Cha, D.; Chou, K. W.; Fischer, A.; Amassian, A.; Asbury, J. B. & Sargent, E. H., Colloidal-quantum-dot photovoltaics using atomic-ligand passivation, Nat Mater, 2011, 10, 765-771 19 Sahu, A.; Kang, M. S.; Kompch, A.; Notthoff, C.; Wills, A. W.; Deng, D.; Winterer, M.; Frisbie, C. D. & Norris, D. J., Electronic Impurity Doping in CdSe Nanocrystals, Nano Letters, 2012, 12, 2587-2594 20 Shim, M. & Guyot-Sionnest, P., n-type colloidal semiconductor nanocrystals, Nature, 2000, 407, 981-983 53 21 Nag, A.; Chung, D. S.; Dolzhnikov, D. S.; Dimitrijevic, N. M.; Chattopadhyay, S.; Shibata, T. & Talapin, D. V., Effect of Metal Ions on Photoluminescence, Charge Transport, Magnetic and Catalytic Properties of All-Inorganic Colloidal Nanocrystals and Nanocrystal Solids, Journal of the American Chemical Society, 2012, 134, 13604-13615 22 Geyer, S. M.; Allen, P. M.; Chang, L.-Y.; Wong, C. R.; Osedach, T. P.; Zhao, N.; Bulovic, V. & Bawendi, M. G., Control of the Carrier Type in InAs Nanocrystal Films by Predeposition Incorporation of Cd, ACS Nano, 2010, 4, 7373-7378 23 Voznyy, 0.; Zhitomirsky, D.; Stadler, P.; Ning, Z.; Hoogland, S. & Sargent, E. H., A Charge-Orbital Balance Picture of Doping in Colloidal Quantum Dot Solids, ACS Nano, 2012, 6, 8448-8455 24 Zhitomirsky, D.; Furukawa, M.; Tang, J.; Stadler, P.; Hoogland, S.; Voznyy, 0.; Liu, H. & Sargent, E. H., N-Type Colloidal-Quantum-Dot Solids for Photovoltaics, Advanced Materials, 2012, 24, 6181-6185 25 Ning, Z.; Zhitomirsky, D.; Adinolfi, V.; Sutherland, B.; Xu, J.; Voznyy, 0.; Maraghechi, P.; Lan, X.; Hoogland, S.; Ren, Y. & Sargent, E. H., Graded Doping for Enhanced Colloidal Quantum Dot Photovoltaics, Advanced Materials, 2013, 25, 1719-1723 26 Oh, S. J.; Berry, N. E.; Choi, J.-H.; Gaulding, E. A.; Paik, T.; Hong, S.-H.; Murray, C. B. & Kagan, C. R., Stoichiometric Control of Lead Chalcogenide Nanocrystal Solids to Enhance Their Electronic and Optoelectronic Device Performance, ACS Nano, 2013, 7, 2413-2421 27 Gregg, B. A., Excitonic Solar Cells, The Journal of Physical Chemistry B, 2003, 107, 4688-4698 28 Thompson, B. & Frdchet, J., Polymer-Fullerene Composite Solar Cells, Angewandte Chemie International Edition, 2008, 47, 58-77 29 Luque, A.; Marti, A. & Stanley, C., Understanding intermediate-band solar cells, Nat 54 Photon, 2012, 6, 146-152 30 Nozik, A. J.; Beard, M. C.; Luther, J. M.; Law, M.; Ellingson, R. J. & Johnson, J. C., Semiconductor Quantum Dots and Quantum Dot Arrays and Applications of Multiple Exciton Generation to Third-Generation Photovoltaic Solar Cells, Chemical Reviews, 2010, 110, 6873-689 31 Ross, R. T. & Nozik, A. J., Efficiency of hot-carrier solar energy converters, Journal of Applied Physics, 1982, 53, 3813-3818 32 Britt, J. & Ferekides, C., Thin-film CdS/CdTe solar cell with 15.8% efficiency, Applied Physics Letters, 1993, 62, 2851-2852 33 Repins, I.; Contreras, M. A.; Egaas, B.; DeHart, C.; Scharf, J.; Perkins, C. L.; To, B. & Noufi, R., 19.9%-efficient ZnO/CdS/CuInGaSe2 solar cell with 81.2% fill factor, Progress in Photovoltaics: Research and Applications, 2008, 16, 235-239 34 Miranda, P. B.; Moses, D. & Heeger, A. J., Ultrafast photogeneration of charged polarons in conjugated polymers, Phys. Rev. B, 2001, 64, 08120135 Arkhipov, V. I. & Bassler, H., Exciton dissociation and charge photogeneration in pristine and doped conjugated polymers, physica status solidi (a), 2004, 201, 1152-1187 55 Chapter 3 Experimental Methods In Chapter 1 and 2, motivational and theoretical background for QD-based photovoltaic devices is presented. In this chapter, detailed description of the main synthesis, fabrication, and characterization techniques is given, while a number of techniques also adopted but less common to this thesis will be introduced thereafter when needed. 3.1 Fabrication of Pbs QD/CdS Bilayer Heterojunction Devices 3.1.1 PbS Colloidal QD Synthesis and Purification Materials: lead acetate trihydrate, oleic acid, 1-octadecene, hexamethyldisilathiane, anhydrous hexane, methanol, ethanol, isopropyl alcohol, butanol, and acetone are purchased and used without purification. Due to the need for large quantity of materials, scale-up synthesis of PbS colloidal QDs, initially developed by a former graduate student, Dr. Scott Geyer, in the group, is adopted in this study. Synthesis was carried out with a Schlenk line. 11.380 g of lead acetate trihydrate, 21 mL of oleic acid, and 300 mL of 1-octadecence was loaded into a IL four-neck flask equipped with a thermal couple, condenser, mechanical stirrer, and heating mantle. The solution was stirred and underwent rough degassing until no vigorous bubbling observed. Then, mild heating at 100 degree Celsius is used to dissolve lead acetate trihydrate and form the lead precursor, lead oleate. Degassing process was continued overnight until 300 millitorr or lower pressure is achieved. 3.15 mL of hexamethyldisilathiane and 150 mL of degassed 1-octadecence were loaded into an injection barrel in a nitrogen glovebox, which was used as the sulfur precursor. The degassed lead precursor was further heated to 120 degree Celsius under nitrogen, followed by swift injection of the sulfur precursor. The solution changed from colorless to black, indicating the formation of PbS QDs. To implement the injection process, an injection gun pressurized at 70 psi was used. It should be noted that the heating apparatus was switched off right before the injection to avoid abrupt heating which could cause non-uniform solution temperature rise and 56 undesired particle size distribution. After injection, the heating mantle was removed to allow cooling of QD solution. After synthesis, the solution was transferred to a vacuumed and sealed flask via a cannula and was pumped into a nitrogen glovebox for initial purification. Appropriate amount of acetone was added to the solution that was then centrifuged at 3900 rpm for 3 minutes to precipitate QDs. The supernatant was decanted and the precipitated QDs were redispersed in hexane at high concentration for the convenience of further purification. The resultant QD hexane solution was stored as a stock solution in a nitrogen glovebox. Before applying to devices, QDs were purified (centrifuged) for two more times with the following recipe. Appropriate amount of ethanol was added to QD solution, accompanied by the addition of butanol with volume equal to one tenth of QD solution to prevent phase separation between ethanol and hexane. The solution mixture was centrifuged at 3900 rpm for 3 minutes to precipitate QDs that were redispersed in hexane and run through 0. 1-micron PTFE syringe filters. The resultant QD solution added with acetone was centrifuged again at 3900 rpm for 3 minutes to precipitate QDs which were redispersed in hexane or octane for dip coating, spin coating, and spray coating as will be discussed in the following Chapter. 3.1.2 Substrate Cleaning Patterned ITO (tin-doped indium oxide)-coated glass substrates purchased from Thin Film Devices Inc. were cleaned by ultrasonication in 2% Micro-90, deionized water, acetone, and boiling isopropyl alcohol for 20 minutes each, followed by five-minute oxygen plasma which removes hydrocarbon contaminants. (The plasma treatment also improves wetting of aqueous solution of PEDOT:PSS (poly(3,4-ethylenedioxythiophene):poly(styrenesulfonate)) used in P3HT (Poly(3-hexylthiophene-2,5-diyl)) nanofiber/CdS QD bulk heterojunction devices as will be introduced later in this chapter.) 3.1.3 (3-mercaptopropyl)trimethoxysilane Substrate Treatment (3-mercaptopropyl)trimethoxysilane (MPTMS) treatment was performed to improve adhesion between QDs and ITO-coated glass substrates. Without MPTMS treatment, PbS QD film was often peeled off during the subsequent postdeposition ligand exchange process. Cleaned substrates were immersed in a jar containing 20 miroliter of MPTMS and 8 mL of toluene in a 57 nitrogen glovebox overnight. Treated substrates were rinsed with isopropyl alcohol in a nitrogen was glovebox and immersed into another jar filled with isopropyl alcohol. The jar with substrates sealed and brought out to the air for 10-minute ultrasonication, after which it was pumped back into a nitrogen glovebox. The substrates were rinsed again with isopropyl alcohol and blew dry with nitrogen. Formation of idealized MPTMS monolayer is illustrated in Figure 3-1 as proposed oxide by other group [1]. In a nutshell, the oxysilane group covalently binds the hydroxylated surface, and the thiol group binds the metal atoms (such as lead atoms, not shown here) in QDs through metal-thiolate bonds. SH I (CH 2)3 I. SH I (CH 2)3 MeO I OMe OMe MeO SH CH 30H Sii Ome OMe OH OH OH OH OH H20 I layer I (CH 2)3 I (CH SH SH SH I I ) 23 H2 0 111 t I OH HO I OH OH O OHOH O I I1 I HO (CH 2 )3 HOI OH O (CH 2 )3 0 I OH OH O OH Figure 3-1. Schematic of idealized MPTMS monolayer formation on hydroxylated oxide substrate surface with the presence of a physically-absorbed water layer. 3.1.4 Deposition of PbS QD Film PbS QD film in PbS QD/CdS bilayer heterojunction devices was deposited via dip coating due to the superior film quality this coating technique can achieve. Purified PbS QDs were redispersed in hexane at a concentration of 10 mg/ml with the addition of 3 droplets of oleic acid. The QD solution was filtered through 0.02-miron AnotopTM syringe filters and contained in a 1 OmL beaker. 0.02% (by volume) 1,2-ethanedithiol (EDT) in acetonitrile and pure acetonitrile was contained in another two beakers separately. MPTMS-treated ITO substrates, held by a steel clamp, were sequentially dipped into the beakers containing QD solution, EDT solution, and acetonitrile, using a multi-vessel dip coater (Nima DC Multi 8, Nima Technology) as shown in Figure 3-2. Specific dip coating parameters are provided in Table 3-1. Because hexane evaporates readily, it was added to QD solution every hour to keep total solution volume constant. After dip coating, QD film on the glass side of ITO substrates were wiped off with a cotton swab moistened with hexane. 58 Figure 3-2. A photograph of the multi-vessel dip coater used to deposit high-quality PbS QD film in this thesis. Beaker Dipping speed (mm/min) immersion time(s QD solution 400 1 30 EDT solution 400 30 0 Acetonitrile 400 3 60 Dyngtm(s Table 3-1. Parameters for PbS QD dip coating. A dip coating cycle consists of: (1) dipping the substrate into 10mg/ml PbS QDs in hexane for 1 second and drying the sample for 30 seconds, (2) dipping the film in 0.02% EDT in acetonitrile for 30 seconds, and (3) rinsing by dipping in pure acetonitrile for 3 seconds, followed by drying for 60 seconds. The above dip coating cycle was repeated to achieve desired PbS QD thicknesses. The dipping speed for all the steps was 400 mm/min. Before and after dipping into QD solution, the samples were dried naturally for 60 and 30 seconds respectively to avoid cross-contamination between QD solution and acetonitrile or EDT solution, which can produce QD agglomerates in solution, resulting in poor film qualities. 3.1.5 Deposition of CdS Thin Film CdS thin film was deposited onto PbS QD-coated ITO substrates via chemical bath deposition in a chemical fume hood. Chemical bath deposition can be regarded as a liquid-phase 59 analogy of chemical vapor deposition. In this section, the mechanism for chemical bath deposition of CdS thin film is first reviewed, followed by the detailed description of the experimental method. Two types of reaction (heterogeneous and homogeneous) occur in chemical bath deposition. The heterogeneous reaction is characterized by the atom-by-atom growth of film on substrates [2], while the homogeneous reaction competes with the heterogeneous reaction and consumes the reactants to form particles in the bulk of the solution [3]. It also has been observed that the predominant homogeneous reaction terminates the heterogeneous reaction, and the film grown by the absorption of massive suspending particles results in porous morphology. Figure 33 shows a rough film grown predominantly by a homogenous reaction. Since a rough film with peaks on the order of micrometer will cause the shorting especially when the overlaying PbS thin film is just hundred-nanometer thick, it is desired to suppress the homogeneous reaction and promote the heterogeneous reaction to obtain smooth CdS film. The following equations list the proposed reactions that occur in the CdS chemical bath deposition [2, 3]: (3-1) NH 3 + H2 0 - NH+ + OH- Cd2+ + 4NH 3 - Cd(NH 3 ) 4 2 + (3-2) Cd2+ + 20H SC(NH 2) 2 + 20H - -3 Cd(OH) 2 (s) (3-3) S2 - + CN 2 H2 + 2H 2 0 (3-4) Cd2+ + S2- -> CdS (s) (3-5) Cd(NH 3) 4 2 + + 20H- + site # [Cd(OH) 2 ]ads + 4NH 3 (3-6) [Cd(OH) 2]ads + SC(NH 2 ) 2 -+ [Cd(SC(NH 2) 2)(OH) 2 ]ads (3-7) [Cd(SC(NH 2 ) 2 )(OH) 2]ads -* CdS + CN 2 H2 + 2H 2 0 + site (3-8) Equations 3-(3-4 and 3-(3-5 are responsible for the homogeneous reaction which generates a large number of CdS particles in the bulk solution and cause the roughness in the resultant thin film. Ammonium hydroxide is added to form cadmium tetraammine complex ions (equation 3(3-2) and prevent the reaction in equation 3-(3-5. Moreover, the reaction in euqation 3- (3-3 also needs to be suppressed because it forms Cd(OH) 2 grains in the bulk solution. By the addition of ammonium hydroxide, cadmium ions are predominantly used up in equation 3-(3-2, and the 60 product of [Cd2+] and [OH] is ensured to be less than the solubility product (1.2 X 10-4) of Cd(OH) 2. Since ammonium hydroxide introduces both NH' and OH (OH~ promotes the unwanted reaction in equation 3- (3-3), ammonium acetate is further added to control the concentration of OH- and stabilize cadmium tetraammine complex ions. On the other hand, equation 3-(3-6 through 3-(3-8 manifest the mechanism of heterogeneous reaction through which smooth and compact films can grow on the substrate surface. This involves reversible adsorption of cadmium hydroxide species on the susbstrate (equation 3-(3-6), formation of surface complex with thiourea (equation 3-(3-7), and formation of CdS with site regeneration (equation 3-(3-8). Figure 3-3. Porous and rough film grown by a homogeneous reaction. In the following, the detailed deposition recipe is presented. 138 mg of cadmium acetate and 308 mg of ammonium acetate were dissolved in 85.9 mL of deionized water in a 150mL beaker at 80 degree Celsius and stirred at 300 rpm. It is noteworthy that because of the slow temperature equilibrium between the reaction solution and the oil bath and because of the sensitivity of the reaction to the temperature, the actual solution temperature was measured. 4.1 mL of ammonium hydroxide (28% NH 3) and 91 mg of thiourea in 10 mL of deionized water were then sequentially added to the 150mL beaker, followed by the immersion of PbS QDcoated ITO substrates which were vertically held by custom-built substrate holders (Figure 3-4) made of Teflon. The reaction was timed since the immersion of substrates. The solution began to 61 turn yellow after 3 minutes, indicating the formation of CdS. The substrates were withdrawn from the solution after 15 minutes and rinsed with deionized water, resulting in 50nm thick CdS film on PbS QD-coated ITO substrates. After naturally drying in the air, the CdS/PbS QD-coated substrates were pumped into an evaporator-integrated glovebox. Figure 3-4. A photograph of the custom-built Teflon substrate holder. 3.1.6 Deposition of Al Electrodes Al electrodes were deposited through a shadow mask via thermal evaporation under a base pressure of 2 x 10-6 torr at a rate of -1 A/s. The deposition thickness was monitored with a crystal monitor, and a typical thickness of 100 nm is used for electrodes. 3.2 Fabrication of P3HT Nanofiber/CdS QD Bulk Heterojunction devices 3.2.1 Growth of P3HT Nanofibers P3HT nanofibers were grown from a solvent-assisted self-assembly method. Regioregular P3HT (Sepiolid P200, Rieke Metals, Inc) was dissolved in 1,2-dichlorobenzene at 10 mg/ml. 5% (by volume) cyclohexanone was added to the P3HT solution, which was then allowed to sit on the shelf without stirring overnight. The appearance of the solution, changing 62 from clear and orange to opaque and purple, indicates the formation of crystalline P3HT nanofibers. 3.2.2 CdS Colloidal QD Synthesis, Purification, and n-Butylamine Ligand Exchange Materials: cadmium oxide, oleic acid, 1-octadecene, sulfur powder, anhydrous hexane, methanol, isopropyl alcohol, butanol, and n-butylamine were purchased and used without purification. Synthesis was carried out with a Schlenk line. 514 mg of cadmium oxide, 20 mL of oleic acid, and 150 mL of 1-octadecence was loaded into a 500mL three-neck flask equipped with a thermal couple, condenser, magnetic stirrer, and heating mantle. The solution was stirred and heated to 250 degree Celsius under nitrogen flow to dissolve cadmium oxide and form the cadmium precursor, cadmium oleate. The appearance of the solution, changing from turbid and brown to clear and colorless, indicated the formation of cadmium oleate. The solution was cooled to 100 degree Celsius for degassing. 64 mg of sulfur powers and 20 mL of 1 -octadecene was loaded into a 40mL septa-capped vial stirred and degassed at 100 degree Celsius, followed by heating to 160 degree Celsius to dissolve sulfur powders and form sulfur precursor. The degassed sulfur precursor was swiftly injected to the degassed cadmium precursor solution at 300 degree Celsius under nitrogen flow. After injection, the solution temperature was plummeted to and maintained at 280 degree Celsius for 30 seconds. Then, the solution was cooled down to room temperature naturally with the heating mantle removed. After synthesis, appropriate amount of isopropyl alcohol was added to the QD solution which was then centrifuged at 6000 rpm for 6 minutes in the air to precipitate QDs. The supernatant was decanted and the precipitated QDs were redispersed in hexane at high concentration for the convenience of further purification. The resultant QD solution in hexane was stored as a stock solution in the air. Before n-butylamine ligand exchange, QDs in hexane were precipitated for two more times following the same procedure as described in the preceding paragraph. The precipitated QD powders were pumped into a nitrogen glovebox, redissolved in n-butylamine, and stirred at room temperature for three days. Small amount of butanol and appropriate amount of methanol were then added to the solution, followed by centrifuging at 3900 rpm for 3 minutes to 63 precipitate QDs, which were redissolved in either octane or 1,2-dicholorobenzene at a concentration of 40 mg/ml. 3.2.3 Blending of P3HT Nanofibers and CdS QDs 1:1 (by volume) of 10mg/ml P3HT nanofiber solution and 40mg/ml CdS QD solution were mixed and stirred at 500 rpm and room temperature for three days. Because of the viscosity of the solution, care should be taken during stirring to minimize the splash of solution onto the vial wall. 3.2.4 Deposition of PEDOT:PSS PEDOT:PSS (Clevios TM P VP CH 8000) was filtered through 0.2-micron PES filters and spun cast on cleaned ITO-coated substrates (using the same cleaning procedure as has been described in section 3.1.2), resulting in 40nm thick film. The spin speed was 5000 rpm with acceleration speed of 6000 rpm/s, and the spin time was 60 seconds. After spin coating, a cleanroom swab moistened with deionized water was used to remove PEDOT:PSS film on the ITO pad close to the substrate edge for electrical contact. Otherwise, the PEDOT:PSS film between the ITO pad and the electrical probe used during the electrical characterization increases resistance, leading to poor photovoltaic performance. Finally, the substrates were baked at 140 degree Celsius for 10 minutes on a hot plate in the air to eliminate residual water in the film and pumped into a nitrogen glovebox where the rest of fabrication process was carried out. 3.2.5 Deposition and Treatment of P3HT Nanofiber/CdS QD Layer 30 microliter of P3HT nanofiber/CdS QD solution was spun coated onto PEDOT:PSS/ITO-coated substrates at 1000 rpm and acceleration of 1000 rpm/s for 60 seconds. After drying naturally, the substrates were dipped into an acetonitrile solution with 0.02% (by volume) EDT for 30 seconds and pure acetonitrile for another 30 seconds. The substrates were then placed on a hot plate with the ramping rate of 450 degree Celsius/minute and baked at 175 degree Celsius for 10 minutes. 64 3.2.6 Deposition of Bathocuproine and Mg:Ag/Ag Layers Deposition of bathocuproine (BCP) and Mg:Ag/Ag layers was performed in the Organic and Nanostructured Electronics Laboratory. Mg was degassed in the thermal evaporator at a base pressure of I x 10~ torr and at the roughly same power as would be used to evaporate Mg. After Mg degassing, a 1 Onm thick BCP layer was deposited at the rate of 0.1 nm/s. Then, 50nm thick of Mg and Ag alloy layer was co-evaporated through a shadow mask at the rate of 0.3 and 0.03 nm/s respectively, followed by the evaporation of 70nm thick Ag layer at the rate of 1.5 nm/s. Figure 3-5 illustrates the overlaid patterns of ITO, BCP, and Mg:Ag/Ag electrodes. ITO BCP Mg:Ag/Ag Probe contact Figure 3-5. Schematic illustration of overlaid patterned ITO, BCP, and Mg:Ag/Ag electrodes. The overlap between ITO and metal electrodes defines the ten devices with an individual active 2 area of 1.24 mm on a half-inch glass substrate. The same pattern was also employed for PbS QD/CdS bilayer heterojunction devices. The pattern was designed by Dr. Alexi Arango. 3.3 Material and Device Characterization 3.3.1 Material Characterization Various microscopy and spectroscopy characterization techniques were performed to analyze the optical property, morphology, chemical composition of the materials synthesized for this study. Optical absorption spectra were obtained using a Cary 5000 UV-vis-NIR spectrophotometer to study the absorption profiles of QD and P3HT in solution and thin film, as 65 well as CdS thin film. Diffuse reflectance of the PbS QD/CdS device was measured with an Ocean Optics ISP-REF Integrating Sphere. Complex refractive indices of PbS QD and CdS thin film were determined using a spectroscopic ellipsometer (J. A. Woollam Co. XLS- 100). Film thicknesses of each layer in PbS QD/CdS and P3HT nanofiber/CdS QD devices were determined using a Veeco Dektak 6M stylus profilometer. The Cross-sectional scanning electron microscopy (SEM) image was taken with a Helios Nanolab 600 dual beam focused ion beam milling system to verify the construction of multilayer thin film devices. The acceleration voltage for electrons and Ga ions were 10 and 30 KV, respectively. The top-view SEM image was acquired with a JEOL 6320FV Field-Emission Highresolution SEM under acceleration voltage of 5 kV to characterize the morphology of CdS thin film. Transmission electron microscopy (TEM) images of drop-cast QDs and P3HT from the sample solution on copper grid (400 mesh) with carbon/formvar support film were obtained with a JEOL 200CX General Purpose TEM to investigate the morphology of QDs and P3HT. Height and phase images of P3HT, PbS QD, and CdS thin film was acquired using a Digital Instruments Dimension 3100 atomic force microscope (AFM) operating in tapping mode to study morphology as well. X-ray photoelectron spectroscopy (XPS) was carried out using a PHI VersaProbe II spectrometer equipped with a monochromated Al Ka X-ray source (1486.6 eV) operating at 50W power and 200ptm spot size to examine local bonding of atom in P3HT/CdS QD blend and elemental composition of PbS QD/ CdS bilayer thin film as a function of depth. 3.3.2 Device Characterization Current-voltage and external quantum efficiency measurement provides simple and decisive characterization to photovoltaic devices and was routinely performed to evaluate and study the device performance. Current-voltage characteristics were acquired using a Keithley 2636 system SourceMeter. During the measurement, samples were placed in a custom-built probe fixture in a nitrogen glovebox and connected to the SourceMeter through a 10-lead cable and a Keithley 7001 switch system, which enabled automatic and individual test of nine devices on the samples. AM 1.5G solar illumination was provided by a Newport 96000 solar simulator and calibrated by a Newport 91150V reference solar cell. 66 The measurement of external quantum efficiency was carried out in the Organic and Nanostructured Electronics Laboratory. The incident light was generated by a quartz tungsten halogen lamp (Oriel Instruments), chopped at 45 Hz by a Stanford Research Systems SR540 Optical Chopper, dispersed by an Acton monochromator, transmitted through an optical fiber, and focused onto sample surfaces with beam size smaller than the device active area. The second-order diffraction allowed by the monochromater was removed, using a series of low-pass filters. The light intensity was calibrated using a Newport silicon photodetector (818-UV) and the photocurrent at a short-circuit condition was measured with a Stanford Research Systems SR830 lock-in amplifier. 3.4 Reference 1 Singh, J. & Whitten, J. E., Adsorption of 3-Mercaptopropyltrimethoxysilane on Silicon Oxide Surfaces and Adsorbate Interaction with Thermally Deposited Gold, The Journal of Physical Chemistry C, 2008, 112, 19088-19096 2 Lincot, D. & Borges, R. 0., Chemical Bath Deposition of Cadmium Sulfide Thin Films. In Situ Growth and Structural Studies by Combined Quartz Crystal Microbalance and Electrochemical Impedance Techniques, Journal of The Electrochemical Society, 1992, 139, 1880-1889 3 Ortega-Borges, R. & Lincot, D., Mechanism of Chemical Bath Deposition of Cadmium Sulfide Thin Films in the Ammonia-Thiourea System: In Situ Kinetic Study and Modelization, Journal of The Electrochemical Society, 1993, 140, 3464-3473 4 Oladeji, I. 0. & Chow, L., Optimization of Chemical Bath Deposited Cadmium Sulfide Thin Films, Journal of The Electrochemical Society, 1997, 144, 2342-2346 67 Chapter 4 PbS QD Schottky Photovoltaic Devices 4.1 Introduction As mentioned in section 1.2, by virtue of their solution processability, near-infrared sensitivity, and band gap tunability, PbS QDs are promising material for the absorber layer in photovoltaic devices. (It is worth noting that although PbSe QDs possess same processing and optical advantages, they have been shown to be less stable than PbS QDs. [1]) It has been previously reported by Jason P. Clifford et al. that Schottky barriers can form at junctions between p-type PbS QDs and shallow-work function metals. Because of the structural simplicity, this thesis began with the fabrication and characterization of PbS QD Schottky photovoltaic devices. 4.2 Deposition of QD Film Due to short minority carrier diffusion length and narrow depletion width, the thickness of PbS QD film is typically limited to few hundreds. However, the need for a thin film imposes challenges on the fabrication of devices with high yield and reproducibility because shorts and shunting paths are easily formed. As a consequence, this thesis began with experimentation and evaluation of a couple of deposition methods, namely spin coating, dip coating, and spray coating, in search of a reliable thin-film coating technique. For each deposition method tested herein, PbS QDs were synthesized and purified per section 3.1.1, and the substrates were cleaned and treated with MPTMS per section 3.1.2 and 3.1.3, respectively. 4.2.1 Spin Coating PbS QD solution (25 mg/ml) in octane was filtered through 0.1-micron PTFE filters and spun cast on ITO-coated glass substrates. The spin speed was 1500 rpm with an acceleration speed of 1500 rpm/s, and the spin time was 60 seconds. EDT treatment was performed by dipping the substrates into an acetonitrile solution with 0.02% (by volume) EDT for 30 seconds 68 PbS QD spin coating and EDT treatment was and pure acetonitrile for another 30 seconds. repeated alternately for 8 times. 4.2.2 Dip Coating PbS QDs were redispersed in hexane at a concentration of 10 mg/ml, filtered through 0.02-miron AnotopTM syringe filters and contained in a IOmL beaker. 0.02% (by volume) EDT solution in acetonitrile and pure acetonitrile solution were contained in another two beakers separately. ITO-coated substrates were sequentially dipped into the beakers containing QD solution for 1 second, EDT solution for 30 seconds, and acetonitrile for 3 seconds. The dip coating and EDT treatment was repeated alternately for 8 times. After dip coating, QD film on the glass side of ITO substrates were wiped off with a cotton swab moistened with hexane. 4.2.3 Spray Coating A commercially available handheld airbrush (Badger Air-Brush Co 200-1 Siphon Feed 200NH Airbrush) powered by compressed nitrogen at 8 psi was used to spray QD thin film. Figure 4-1 shows the schematic representation of the spray coating setup employed in this study. PbS QDs were dissolved in octane at a concentration of 5 mg/ml and contained in a jar connected to the airbrush. The flow of compressed nitrogen gas through a venturi creates a local reduction in gas pressure, sucking QD solution from the jar. The high velocity of nitrogen gas atomizes QD solution into fine droplets and carries the droplets onto the substrate surface. The spray of QD solution was raster scanned on the substrate surface to produce a thin layer of QD film, followed by EDT treatment and acetonitrile rinse as done with spin or dip coating. The spray coating and EDT treatment was repeated alternately for 8 times with the substrate rotated 90 degrees between each iteration. 69 Nitrogen gas pressure Solvent vapor pressure Solution feed rate Tip-to-substrate distance Figure 4-1. A schematic representation of a spray coating setup composed of a handheld airbrush, a gas regulator, and a nitrogen cylinder. Relevant processing parameters are also shown. To achieve a good film quality, several spray coating parameters have to be diligently controlled. Major solvent evaporated before the droplet arriving at the substrate surface would lead to poor QD adsorption on the substrate surface and therefore should be avoided. For this concern, the relevant parameters are solvent vapor pressure (or the solvent evaporation rate) and the distance over which the droplets need to travel before arriving at the substrate surface (called tip-to-substrate distance). On the contrast, once the droplets arrive at the substrate surface, rapid solvent evaporation is preferred to minimize the migration of QDs to the circumference of the droplets, the so-called coffee stain effect, which would cause non-uniform distribution of QDs on the substrate surface. For this concern, the relevant parameters are the solvent evaporation rate and the solution feeding rate. Additionally, over high speed of carrier gas (nitrogen gas) would smash the droplets onto the substrate surface, further spreading out QDs, and hence should be avoided as well. For this concern, the relevant parameters are the pressure of nitrogen gas and the tip-to-substrate distance. 4.3 Film Morphology and Device Characteristics Optical micrographs of PbS QD film deposited via spin coating, dip coating, and spray coating are shown in Figure 4-2. QD film deposited via spin coating and dip coating were 70 generally very smooth, although spin-coated film did exhibit more circular or comet-like defects. AFM images (Figure 4-3) of spin-coated and dip-coated film confirm that the sub-micron morphology of this film indeed is very similar to each other. Root means squared roughness of the spin coated and the dip coated film is 3.9 nm and 5.1 nm, respectively. On the contrast, spray coated film is much rougher as can been seen in Figure 4-4, resulting in root mean squared roughness of 72.8 nm. The thicknesses of spin-coated, dip-coated, and spray-coated QD film are 150 nm, 170 nm, and 180 nm, respectively. 100 m Figure 4-2. Optical micrographs of PbS QD film deposited via (a) spin coating, (b) dip coating, and (c) spray coating. 35.0 nm 5pVm r5Pm Figure 4-3. AFM images of PbS QD film deposited via (a) spin coating and (b) dip coating. 71 A 6000 6888.5 A 5000 4000 3000 399 pm 2000 1000 400 PM 0 Figure 4-4. A three-dimensional profilometry image of spray-coated PbS QD film. Figure 4-5 presents the J-V characteristics of AM 1.5G illuminated Schottky devices with PbS QD film deposited via spin coating, dip coating, and spray coating. All of these devices demonstrated similar J-V characteristics. The smaller shunt resistance and lower fill factor of the spray-coated device may be attributed to its higher film roughness and thus higher device leakage current. 2 U E 0- 2. 3 4. Spin Coating -0.1 0.0 0.1 0.2 0.3 -4 - 0.4 -0.1 0.0 ' 0.1 Dip Coating 'I 0.2 0.3 ~4 ~ 0.4 -0.1 Spray Coating 0.0 0.1 0.2 0.3 0.4 Voltage (V) Figure 4-5. J-V characteristics of PbS QD Schottky devices under simulated AM1.5G illumination. PbS QD film in these devices was deposited via (a) spin coating, (b) dip coating, and (c) spray coating. Consequently, film morphology and J-V characteristics suggest that spray coating would be a more reliable deposition technique when a thick film, rather than a thin film, is required. It should be also noted that spray coating wastes less QD solution than spin coating [3] and spray coating [4] do and thus is a promising technique when thick and large-area deposition is 72 necessary in terms of material saving. However, spin coating and dip coating are highly recommended when a smooth film is needed. In a film-quality-demanding case, such as the fabrication of PbS QD/CdS bilayer devices, dip coating has to be adopted to deposit a nearly- defect-free film. 4.4 Conclusion In summary, while spray coating can be a viable and cost-effective technology for the deposition of large-area and thick QD film, dip coating provides the best film quality (in terms of smoothness and reduced number of defects) among the three coating technique examined in this thesis and hence is preferred when film quality is the top priority. Overall, the Schottky devices exhibit poor PCE. It has been known that a Schottky diode (a majority-carrier device) typically shows a higher dark current and thus a lower turn-on voltage and Vc than a pn diode (a minority-carrier device) does [5]. Additionally, Fermi-level pinning semiconductor interface is also responsible for low V, at the metal and in a Schottky device [6]. Therefore, the bulk of this thesis work was to develop a PbS QD solar cell based on a pn junction which will be detailed in Chapter 5. 4.5 Reference I Luther, J. M.; Law, M.; Beard, M. C.; Song, Q.; Reese, M. 0.; Ellingson, R. J. & Nozik, A. J., Schottky Solar Cells Based on Colloidal Nanocrystal Films, Nano Letters, 2008, 8, 3488-3492 2 Clifford, J. P.; Johnston, K. W.; Levina, L. & Sargent, E. H., Schottky barriers to colloidal quantum dot films, Applied Physics Letters, 2007, 91, 253117 3 The majority of the solution is spun out of substrates and wasted during spin coating. 4 QD solution is contaminated due to repeated substrate immersion and necessitates routine filtration which sacrifices some of the solution. In addition, QDs coated on the back side of the substrate need to be removed and are wasted, too. 5 Green, M. A., Solar Cells, 1998, 176 73 6 Luther, J. M.; Law, M.; Beard, M. C.; Song, Q.; Reese, M. 0.; Ellingson, R. J. & Nozik, A. J., Schottky Solar Cells Based on Colloidal Nanocrystal Films, Nano Letters, 2008, 8, 3488-3492 74 Chapter 5 PbS QD/CdS Bilayer Heterojunction Photovoltaic Devices 5.1 Introduction Low-temperature solution-processed solar cells incorporating organic semiconductors [1, 2] and colloidal quantum dots (QDs) [3] are a potential alternative to conventional solar cells fabricated via vacuum or high-temperature sintering processes for large-area, high-throughput, and low-cost manufacturing. While there has been significant effort to synthesize low-band gap organic semiconducting molecules for photovoltaic applications [4], PbS and PbSe QDs can already be tuned across the near-infrared spectrum from k=2200nm to k=600nm wavelength [5, 6, 7], encompassing the range of ideal band gaps for optimum efficiency in both single- and multi-junction solar cells [8, 9]. For example, tunable band structures facilitate implementation of cascaded band gaps in the multi-junction stack geometry, which can optimize the light capture of the solar cell stack. Since the report of the first PbS QD Schottky junction solar cell in 2008 [10] there have photovoltaics. been steady improvements in the power conversion The highest currently-reported efficiency efficiency of QD of 7.0% was achieved through passivation of the PbS QDs with halide anions and organic thiol, incorporating TiO 2 annealed at 520"C [11]. Recently, the low-temperature fabrication of PbS/Bi 2 S 3 QD bulk heterojunction and PbS QD pn junction solar cells exhibiting efficiencies of 4.9% [12] and 5.4% [13], respectively, were reported. Here, we explore an alternative heterojunction solar cell based on PbS QDs and a low-temperature solution-processed CdS n-type inorganic semiconductor window layer adapted from traditional CdTe photovoltaic technology. Our fabrication process utilizes the chemical bath deposition technique, which has been shown to be a low-temperature, solution-phase, and scalable method for the fabrication of numerous inorganic semiconductors [14, 15]. For example, chemical bath deposition is used to fabricate CdS thin films as n-type window layers for conventional CdTe and CIGS solar cells with efficiencies as high as 15.8% [16, 17]. It has been shown that the superior performance of CdTe solar cells coated with CdS is due to field quenching, attributed to the excitation of trapped holes and recombination of electron-hole pairs in the CdS [18]. Analogous to this work, in the 75 present study we chose chemical-bath-deposited CdS to form a rectifying heterojunction with PbS QDs, enabling us to assess the compatibility of this deposition process and PbS QDs. 5.2 Device Structure and Fabrication Figure 5-1 (a, b) shows the device architecture and the band diagram of our PbS QD/CdS& bilayer heterojunction solar cells. PbS QDs of a wide band gap (-1.5 eV, as determined from the first absorption excitonic peak at X=850 nm shown in Figure 5-2) are used to ensure energetically-favorable electron transfer from PbS QDs to CdS and to reduce the interfacial recombination current [19]. Energy levels of the constituent materials are taken from the literature [20, 21]. Light is illuminated on the heterojunction from the side of the PbS where excited states are mainly generated in the smaller-energy-band-gap PbS QDs, and the larger gap CdS can act as an optical spacer. Aspects of optical design are subsequently explored with the use of optical simulations. (a) (b) CdS 4.2 eV Al ITO PbS QDs ITO-Glass- Al 4.2 eV PbS 4.8 eVI CdS 6.6 6V Figure 5-1. (a) Schematic of the PbS QD/CdS heterojunction device fabricated in this study. (b) Energy band diagram showing band edges of isolated 1.5eV PbS QDs and CdS along with electrode work functions. 76 ,-2.0 . , PbS QDs M1.0 U C 20.50 <0. 300 400 500 600 700 800 900 1000 Wavelength (nm) Figure 5-2. Absorbance spectrum of PbS QDs in solution used in this work. Material synthesis and device fabrication have been detailed in Chapter 3 and is briefly summarized here. PbS QDs were prepared by a modified recipe reported by Hines et al. [5], then deposited onto ITO-coated glass substrates using a layer-by-layer dip-coating method [22] employing EDT ligand exchange to increase carrier mobility and insolubilize the QDs. The thickness of the PbS QD layer was determined by the number of dip-coating cycles and ranged from 40 nm to 280 nm as measured with a profilometer. CdS thin films were deposited onto PbS QDs at 80 0C in an aqueous solution. The SEM image in Figure 5-3 displays the polycrystalline nature of these chemical-bath-deposited CdS thin films with grain sizes around 200 nm. The electrical conductivity measurement provided in Figure 5-4 reveals that CdS thin films are five orders of magnitude more conductive than PbS QD thin films of similar thickness. The relatively high conductivity of as-synthesized CdS thin films guarantees that the series resistance added to the PbS QD devices is negligible, thus eliminating the need for further vacuum or annealing processes. The cross-sectional SEM image and depth-profile XPS presented in Figure 5-5 confirm the layer structure of the PbS QD/CdS heterojunction devices. 77 Figure 5-3. SEM image of the CdS thin film on glass. 1E-5 1E-6 1E-7 1E-8 C 1E-9 1E-10 -+-PbS + CdS 1E-11 1E-12 -1.0 -0.5 I 0.0 0.5 1.0 Voltage (V) Figure 5-4. Current-voltage characteristics of PbS QD and CdS thin films. The electrical conductivities of PbS QDs and CdS were measured using a planar interdigitated configuration. For PbS QDs, the channel length is 6 ptm, the channel width is 1.95 cm, and the film thickness is 100 nm. For CdS, the channel length is 120 pm, the channel width is 0.75 cm, and the film thickness is 120 nm. Consequently, the conductivities of PbS QDs and CdS are ~-2 x10-8 and -1 x 10-3 S/cm, respectively. 78 (a) (b) 500 450 -Cd3p3/2 01s -- Pb4f -- S2s 400 350 Pt 300 Al2s *~250 200 CdS S150- PbS QDs100 50 ITO0 .l..- 0 10 20 30 40 50 60 70 Sputter Time (min) Figure 5-5. (a) Cross-sectional SEM image of the device. Platinum was deposited on top of the device before focused ion beam milling to protect the underlying layers. (b) XPS depth profile of a PbS QD/CdS device without Al electrodes. 5.3 Protocols for the Fabrication of Reproducible Heterojunction Devices During the course of study, device yield and reproducibility had once been an issue such that drawing a convincing and reliable conclusion had been proven difficult. J-V characteristics of nine PbS QD/CdS devices made on the same substrate were plotted in Figure 5-6 to illustrate this problem. The representative optical microscopy image (Figure 5-7) reveals that the film in these devices was characterized by pinholes. Pinholes can create shunting paths and shorts in the devices, leading to a leaking diode with I-V characteristics akin to those shown in Figure 5-6. Therefore, we attempted to find out the origin of pinhole formation. 79 20NJ' E 15 - Dark -- AM1.5 G 10- E VLb E 0 U -15 -1 -0.8 -0.6 -0.4 -0.2 0 0.2 0.4 0.6 0.8 1 Voltage (V) Figure 5-6. J-V characteristics of nine PbS QD/CdS heterojunction devices fabricated on the same substrate and measured under simulated AMI.5G solar illumination. The scattered J-V characteristics manifest the device irreproducibility. Figure 5-7. A representative top-view optical micrograph of a finished PbS QD/CdS device. Multiple pinholes can be identified in the image. Because this was a through-thickness pinhole, we were able to concentrate our attention on its formation in either ITO or PbS QD layer. The fact that ITO is mechanically and chemically robust during PbS QD and CdS deposition processes allow us further exclude the possibility of pinhole formation in the ITO layer. Therefore, we could focus solely on the PbS QD layer. Intuitively, the question was whether the pinhole developed during or after PbS QD deposition (i.e. during subsequent fabrication processes). From the optical micrographs shown in Figure 4-2 and Figure 5-8(a), we know that the dip-coated QD film was nearly defect-free: there 80 particles but no pinhole observed. In addition, the Schottky devices fabricated with the same recipe in the same batch exhibits reproducible J-V characteristics as shown in Figure 5-8(b). This observation infers that the pinhole did not develop during PbS QD were only few deposition. (a) (b) 20 ' 15 E 10 -- Dark - AM1.5 G E 54-J o 0 -5- 4-J -10 - 15 -' 91 -Ov8 0 0.2 0.4 0.6 0.8 Voltage (V) -O.6 -0.4 -0.2 1 Figure 5-8. (a) A representative top-view optical micrograph of finished PbS QD Schottky devices. No pinhole is identified in this image. (b) J-V characteristics of nine PbS QD Schottky devices fabricated on the same substrate and measured under simulated AM1.5G solar illumination show good reproducibility. After carefully comparing the PbS film before (Figure 5-9(a)) and after CdS deposition (Figure 5-9(b) and (c)), we found that although some particles stayed in the film, some were peeled off from the film, forming pinholes. Therefore, we attempted to tackle the pinholes by minimizing their predecessors-particles. 81 Figure 5-9. Representative optical micrographs of (a) as-deposited PbS QD film, (b) PbS QD film with a CdS overlayer and with no pinhole, and (c) PbS QD film with a CdS overlayer and with a pinhole. The pervasive submicron dark spots in (b) and (c) are the CdS crystallites which were initially formed in the bulk of the chemical bath solution and subsequently adsorbed and embedded in CdS film. After examining the fabrication process, we identified three possible routes for the formation of particle defects in the PbS QD film and took the according measures to prevent. First, the dust particles or other surface contaminants, if not eliminated during the substrate cleaning procedure, on the ITO-coated glass substrates would be covered by PbS film and appear as particle defects in the film. The optical micrograph (Figure 5-10) of the substrate undergoing one cycle of the substrate cleaning process detailed in Chapter 3 shows the existence of dust particles/contaminants. Therefore, to ensure particle- and contaminant-free surfaces, the substrates underwent multiple cleaning cycles and were inspected under an optical microscope prior to subsequent fabrication processes. Figure 5-10. An optical micrograph of an ITO-coated glass substrate cleaned with the procedure described in Chapter 3 still shows dust particles/contaminants (highlighted with the red circles) adsorbed on the surface. 82 Second, we found that thoroughly-cleaned substrates were contaminated again during the MPTMS treatment as shown in Figure 5-11. It was reported by Minghui Hu et al. that excess water in solvents used for the treatment accelerates self-polymerization of MPTMPS molecules, which not only become the sources of particle defects but also reduce the binding capability of the adhesion layer formed by these molecules [23]. To avoid the self-polymerization, the MPTMS treatment, including subsequent rinsing steps, was performed in a nitrogen dry glovebox. 500 _____M gim Figure 5-11. Optical micrographs of ITO-coated glass substrates (a) before and (b) after MPTMS treatment. The red circles in (b) highlight the surface contaminants. Last, QDs with poor colloidal stability aggregates and grows into big particles. Figure 512 presents an optical micrograph of the PbS QD film deposited from a solution with poor colloidal stability. It can be seen that the film is scattered with small spots. Figure 5-13 demonstrates that essentially all of the QDs in the solution added with excess amount of ethanol were precipitated at the bottom of the tube after centrifugation. Excess alcohol, such as ethanol, added in the purification process is linked to loss of surface ligands [24], which can cause the aggregation of QDs. Therefore, to preserve the native ligands on QD surfaces and thus maintain good colloidal stability, the amount of ethanol and acetone added to the QD solution during purification process described in Chapter 3 was reduced, and two drops of native ligands (oleic acid) were added to the final purified QD solution (~10 mL). Additionally, before and during the film deposition processes, the QD solution was filtered with 0.02-miron AnotopTM syringe filters to remove aggregated QDs. 83 500 pm Figure 5-12. An optical micrograph of PbS QD film deposited from a solution of poor colloidal stability. Multiple particle defects can be identified in the image. Figure 5-13. Photographs of centrifuged PbS QD solution added with (a) excess and (b) reduced amount of ethanol. Complete purification process should be referred to Chapter 3. The supernatant in the solution with excess amount of ethanol is much clearer than that with reduced amount of ethanol, suggesting that most of the QDs was precipitated out during centrifugation in the solution with excess amount of ethanol. After the implementation of the aforementioned measures, nearly-particle-free and pinhole-free devices as shown in Figure 5-14(b) have been routinely made. The current densityvoltage characteristics (Figure 5-15) of the nine PbS QD/CdS devices fabricated according to the newly-implemented measures on the same substrates also exhibit good reproducibility. 84 1 500 4m 1500 Figure 5-14. Optical micrographs of the PbS QD/CdS devices fabricated (a) without and (b) with additional measures to prevent pinhole formation. The red circles in (a) highlight the particle defects/pinholes. The blurred black spots in both images were the residual of the QD film on the back side of the substrates. During a dip-coating process, both sides of the substrates were coated, and the film on the back side was wiped off before the measurement. Therefore, these blurred spots are not in the device and can be ignored. 30 r4 E 20 U -Dark -AM1.5 G E 10 U -20 -30 -1 -0.8 -0.6 -0.4 -0.2 0 0.2 0.4 0.6 0.8 Voltage (V) 1 Figure 5-15. J-V characteristics of nine PbS QD/CdS heterojunction devices fabricated with the newly-implemented experimental methods on the same substrate and measured under simulated AMI.5G solar illumination. 85 5.4 The Comparison between Schottky and PN Heterojunction Devices The advantages of utilizing pn heterojunctions over Schottky junctions for photovoltaics have been discussed by Pattantyus-Abraham et al. [25] In particular, heterojunctions can minimize the Fermi-level pinning that frequently occurs at metal-semiconductor interfaces, which limits the Schottky barrier height and V0 c. In addition, wide band discontinuities at pn heterojunctions can suppress the contribution of electron or hole injection to the dark current and benefit both the FF and the Vc. Figure 5-16 displays the representative dark and light J-V characteristics of PbS QD Schottky and PbS QD/CdS bilayer heterojunction devices under AM1.5G illumination. Both of these devices consist of a 200nm thick layer of PbS QDs, prepared using the same EDT ligand exchange procedure, and sandwiched between an ITO bottom anode and an Al top cathode. The Vc, J, and PCE of heterojunction devices unambiguously outperform those of Schottky devices, indicating the effectiveness of using chemical-bath-deposited CdS as the n-type semiconductor in heterojunction devices. 10 r4i E 5- E0 -5- -- -15- .2 -0.1 Schottky, dark Schottky, AM1.5G Heterojun., dark Heterojun., AM1.5G 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 Voltage (V) Figure 5-16. J-V characteristics of typical PbS QD Schottky (red) and PbS QD/CdS heterojunction (blue) devices measured in the dark and under AMi.5G simulated solar illumination. The applicability of CdS as an n-type material in pn heterojunctions is further demonstrated by the inverted solar cell shown in Figure 5-17. The inverted CdS/PbS QD solar cell with symmetric ITO top and bottom contacts displays inverted polarity compared with the 86 normal PbS QD/CdS, PbS QD Schottky, or ITO/PbS/ITO solar cells and an apparent V, of -0.32 V, indicating that the built-in potential of the CdS/PbS QD heterojunction is greater than the top Schottky barrier that has been reported in several paper [22, 26, 27]. -N 2.0 E E - 1.5 1.0 0.5 0.0 In -0.5 0 -1.0 -AM1.5G -1.5 -dark U -2n -0.5-0.4-0.3-0.2-0.1 0.0 0.1 0.2 0.3 0.4 0.5 Voltage (V) Figure 5-17. J-V characteristics of an inverted CdS/PbS QD heterojunction solar cell measured in the dark and under AM1.5G simulated solar illumination. To verify the origin of the device rectifying behavior, symmetric top and bottom ITO contacts were employed. Due to some spikes of chemical-bath-deposited CdS thin films, a thick (310 nm) PbS QD layer was used in this inverted device to avoid shorting, which may result in a worse photovoltaic performance than a normal device. 5.5 PbS QD Layer Thickness-Dependent J-V Characteristics Because the performance of thin film solar cells is highly dependent on the active layer thickness due to a tradeoff between light absorption and carrier extraction, we studied the effect of the PbS QD layer thickness on the J-V characteristics to optimize device performance. Representative J-V curves for PbS QD/CdS heterojunction solar cells with various PbS QD thicknesses (40, 100, 160, 220, and 280 nm) are plotted in Figure 5-18(a), and the photovoltaic parameters are summarized in Figure 5-18(b). A decrease of V" with increasing thickness of PbS QD film is observed, which is attributed to a reverse Schottky contact (non-ohmic contact) that emerges at the anode, impeding electrical conduction and acting as an opposing diode in series, thus reducing the voltage [26, 27]. When the thickness is small, the built-in potential from the depletion region is sufficient to assist carriers tunneling through this non-ohmic contact. While when the thickness is large, this non-ohmic contact becomes important. Nevertheless, the 87 PCE initially increases with PbS QD layer thickness to a maximum at 160 nm, then decreases with greater PbS QD layer thicknesses. With a 160 nm thick PbS QD layer, we achieved an average J,, of 15.3 ± 1.1 mA/cm2, Voc of 0.54 ± 0.02 V, FF of 0.42 ± 0.01, and PCE of 3.5 ± 0.3%. The trend in PCE is primarily correlated with that of the Js, as illustrated in Figure 5-18. We therefore investigated the origin of the dependence of Js, on PbS QD thickness. 88 (a) I 10 _ E 01 -10 -15 -20 -0.2 -0.1 0 0. 1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 Voltage (V) (b) E E 18 0.5 15- 0.4 g 2-0.3 L 6 0.1 S3. 00.70.6 0.0 4.0 3.5 3.0 * 0.5- 0.4. 2.5 u-3 2.0 1.5. 0.3 > 0.20.1 0.0 0 0.2 1.0. 50 100 150 200 250 300 PbS Thickness (nm) 0.5 0.0 0 50 100 150 200 250 300 PbS Thickness (nm) Figure 5-18. (a) J-V characteristics of representative PbS QD/CdS heterojunction solar cells with varying PbS QD layer thicknesses. (b) Performance characteristics of the above devices. The error bars are evaluated from the standard deviation of 18 devices fabricated with the same condition for each thickness of the PbS QD film. 5.6 Modeling of PbS QD Layer Thickness-Dependent Jc The generation of photocurrent in a pn junction solar cell can be described in terms of two fundamental processes [28]. The first process involves the absorption of incident photons 89 and the creation of electron-hole pairs. Since the constituent materials and fabrication processes are identical for the devices of different thickness, any variation in the generation efficiency of electron-hole pairs following photon absorption can be neglected. Consequently, the first process is determined only by the absorption of incident photons where it is well-known that the optical electric field (and correspondingly the generation rate) inside multi-layered thin films with reflective metal contacts can be modulated dramatically as a result of optical interference between incident and reflected light [29, 30, 31]. In order to understand the spatial distribution of generation rates, we employed an optical model as previously described by Pettersson et al. [32] where the complex indices of refraction (reported in Figure 5-19) and layer thicknesses of the materials were determined by spectroscopic ellipsometry and profilometry. 2.5 2 1.5 - CdS n --- CdSk 1 .5 0 -- PbS n PbS k 0-- 400 500 600 700 800 900 1000 Wavelength (nm) Figure 5-19. Complex refractive indices of PbS QDs and CdS. The model assumes smooth interfaces between each layer, which is justified in this work. Figure 5-20 shows the AFM surface images of PbS QD and CdS thin film. Although there are some spikes observed on the CdS surface, the root mean squared roughness of CdS is ~5 nm, which suggests most of the interface in the device is smooth. The reflectance measurement of the full PbS QD/CdS device (without Al electrode) carried out with an integrating sphere also suggests only a small diffuse reflectance across the visible region (Figure 5-21) where most of photons are absorbed in the device. In addition, the relatively good agreement between the EQE and modeled absorption spectra at low PbS thicknesses shown in Figure 5-22 indicates interface scattering is not a significant effect. Therefore, the optical modeling was simplified without the inclusion of interface roughness. 90 (a) (b) 15.0 nm 50.0 nm I I 0.0 Height 1.5 um 0.0 Height 1.5 uM Figure 5-20. AFM surface images of (a) the PbS QD thin film and (b) the CdS thin film. 100 . . . . . . . . . . . . . 90 U a In 80 70 60 50 40 30 20 10 0 300 400 500 600 700 800 900 1000 Wavelength (nm) Figure 5-21. Diffuse reflectance spectrum of a full PbS QD/CdS device (without Al electrode). 91 (a) .100 90 $ ., 80 -r- 70 . 60 0 50 c 40 .2 30 t 20 L10 0 (b) absorptionEQE - -- (e) 100 90 80 70 oc 4.OR60 0 50 C 40 0 30 U 20 Cu 10 U0 4-1 100 90 80 70 60 O.. -c 40 50 C 40 30 20 U 10 absorption EQE C 40 .0 30 S20 IL 0 400 500 600 700 800 900 1000 400 500 600 700 800 900 1000 Wavelength (nm) (c) .100 # 90 - 80 70 60 (d) absorption- - -- EQE 400 500 600 700 800 900 1000 --6 7 '100 *90 +j 80 -r- 70 60 -O o50 c 40 . 30 U 20 10 0 absorption 400 500 600 700 800 900 1000 Wavelength (nm) Wavelength (nm) - Wavelength (nm) absorption EQE 400 500 600 700 800 900 1000 Wavelength (nm) Figure 5-22. Measured EQE (red curves) and modeled active-layer absorption spectra (black curves) of PbS QD/CdS devices with varying PbS QD thicknesses: (a) 40 nm, (b) 100 nm, (c) 160 nm, (d) 220 nm, and (e) 280 nm. In Figure 5-22, the EQE spectra roughly match the absorption spectra for (a), (b), and (c). The discrepancy seen in (a), (b), and (c) may be due to weak diffusive scattering [33] resulting from slight interface roughness but more likely stems from slight inaccuracies in the thickness of each layer and variations in refractive indices which can sensitively impact the optical 92 interference and cause shifts in the optical spectra; this conclusion is supported by the observation that all the absorption spectra simply appear to be slightly shifted from the measured EQE, but have the same spectral features (peaks, valleys, and tails). The large discrepancy in the short-wavelength range seen in (d) and (e) is attributed to poor carrier extraction for those carriers generated by short-wavelength photons since these photons are absorbed further away from the PbS QD/CdS junction (generating a position dependent IQE). The reason this wavelength (position) dependent IQE is not seen in (a)-(c) is because the thicknesses are smaller than the depletion width. The simulated wavelength- and distance-dependent photon absorption rates under AMl.5G solar photon flux for the complete device with various PbS QD thicknesses are illustrated in Figure 5-23. For ease of visualization, we also integrate the absorption rate with respect to wavelength and plot the total distance-dependent exciton generation rate in the PbS QD layers in Figure 5-24. It is found that the generation rate (or, correspondingly, the rate of photon absorption) does not follow a simple exponential decay with thickness; instead, a local maximum occurs near the interface between PbS QDs and CdS regardless of PbS QD thickness. To elucidate whether such an generation pattern would favor a particular thickness of PbS QDs and because carrier separation is supposed to take place at the heterojunction interface, we integrate the full-spectrum generation rate displayed in Figure 5-24 with respect to the distance from the PbS QD/CdS interface to obtain the integrated exciton generation rate, as depicted in Figure 5-25. Integrated to a given distance from the interface, the generation rates are smaller for thicker PbS QD films since more of the light is absorbed near the incident interface close to the ITO. With this simulation result, we find that even though there is an optical cavity effect in our devices, the thinner devices still have more excited states generated close to the PbS QD/CdS interface depletion region than in the thicker devices. 93 (a) Al PbS CdS ITO x 1019 (b) x 109 6 6 5 E E 4 4 3 C (U CU 0M 2 2 Cu 1 1000M 100 200 Distance (nm) 0 (C) PbS ITO CdS 300 0 0 100 200 0 300 Distance (nm) Al X1o 9 (d) x 101 5 4 E E 4 3 4-J to a 0j W) 2 2 1 0 100 200 300 400 1 0 0 Distance (nm) (e) 100 200 300 400 500 -0 Distance (nm) ,119 x 101, 5 E 4 -C C 2 0M 1 0 100 200 300 400 500 0 Distance (nm) Figure 5-23. Modeled photon absorption rate (1/sec-cm 3 ) for PbS QD/CdS heterojunction devices with varying PbS QD thicknesses: (a) 40 nm, (b) 100 nm, (c) 160 nm, (d) 220 nm, and (e) 280 nm. The x- axis (distance) is measured from the ITO surface where light is incident. The horizontal stripes seen in the simulation plots result from the use of AM1.5G solar spectrum as a light source for the calculated absorption/generation rate. There are multiple dips in the solar spectrum due to light absorption by 03, 02, H2 0, and CO 2 . 94 E 2 40 nm in 0 1 100 nm 0 ------------------------------------ --CDi 0-16 lnm C 1220 01 nm L1 280 nm Co0 CU 0 50 100 150 200 250 300 Distance (nm) Figure 5-24. The modeled total generation rates in PbS QD layers of different thicknesses (offset for clarity). E U 14 Xx10 - 40nm 12 100nm -- 160nm 10 220nm a8 -280nm i8, u 4 24-0 0 50 100 150 200 250 300 Distance from the PbS/CdS Interface (nm) Figure 5-25. The integrated generation rate (1/sec-cm 2 ) as a function of the distance from the PbS QD/CdS interface for the above PbS QD thicknesses. The second process involved in photocurrent production is the extraction of photogenerated carriers. As mentioned above, the carrier extraction efficiency is described by the sum of excition/carrier diffusion length and depletion width relative to PbS QD thicknesses and can be estimated by comparing the device data in Figure 5-18 with the optical modeling results 95 in Figure 5-25. The devices with 40nm and 220nm thick PbS QD layers exhibit similar Jc as shown in Figure 5-18. Consequently, the amount of carriers extracted must be the same for the two devices. Since 40 nm is rather thin compared to the convoluted length of typical depletion width and diffusion length reported for PbS QDs [19, 27], we can assume that all of the carriers generated in the device are extracted. This assumption is supported by the high internal quantum efficiency (IQE) of 93% calculated from the ratio of measured to modeled Jsc (from the modeled total generation rate and assuming 100% IQE) and the good agreement between the measured and modeled J,, (assuming 100% IQE) integrated from the external quantum efficiency (EQE) spectra at low PbS QD thicknesses as illustrated in Figure 5-26. From Figure 5-25, the integrated rate of carrier collection of the 220 nm thick device is seen to match that of the 40 nm thick device at a distance roughly 170 nm into the PbS film from the PbS/CdS interface; this distance constitutes an estimate of the convolution of the diffusion length and the depletion width within the 220 nm thick PbS QD layer. 20 . . . . 4-'15 E E 10. modeled -- measured 150 200 250 300 * 0 50 100 PbS QD Thickness (nm) Figure 5-26. J,, integrated from the modeled (black curve, assuming 100% IQE) and measured (red squares) EQE spectra. 5.7 Capacitance Measurement The depletion width can also be determined from a measurement of the device capacitance. In order to obtain the depletion width in the PbS QD layer in our full device, we measured the capacitance of PbS QD/CdS devices with various PbS QD thicknesses under zerobias conditions [34, 35] rather than implementing Mott-Schottky analysis in which a Schottky junction or a pn junction with n or p much more heavily-doped than p or n must be employed [36]. The capacitance is expected to decrease with increasing thickness as long as the thickness is 96 smaller than the depletion width, in which case the fully-depleted diode resembles a parallelplate capacitor with a dielectric thickness equal to the depletion width. However, when the thickness exceeds the depletion width, the capacitance should be insensitive to the change of thickness. The measured and modeled capacitances are plotted in Figure 5-27. 5 . m Measured A 4 LLA o Modeled lower bound A Modeled upper bound A AU U r2 4-A 3 .- A CU' U0~ 0 50 100 150 200 250 300 PbS Thickness (nm) Figure 5-27. Measured (black square) and modeled (red circle for the lower bounds; blue triangle for the upper bounds) device capacitance at zero-bias as a function of PbS QD thicknesses. The lower bounds were calculated, assuming that both the PbS QD and CdS layer are fully-depleted; while the upper bounds were calculated, assuming that only the PbS QD layer is fully-depleted and none of CdS layer is depleted. The capacitances of PbS QD/CdS heterojunction devices were modeled based on simple parallel-plate capacitors in series, in which case the capacitance C is given by C= -CPbS = CCdS (PbS-E' dPbS) + ECdS Eo dCdS) (5-1) where Cpbs is the capacitance contributed from the PbS QD layer, Ccds is the capacitance contributed from the CdS layer, EPbs is the dielectric constant of PbS QDs (= 18) [34], Ccds is the dielectric constant of CdS (= 8.9) [37], so is the permittivity of free space, A is the device area ( = 1.24 mm 2 ), dPbs is the depletion width in the PbS QD layer, and dcds is the depletion width in the CdS layer ( = 50 nm for the lower bound; = 0 nm for the upper bound). The modeled lower bound is calculated with the assumption that both the PbS QD and CdS layers are fully depleted, while the modeled upper bound assumes that the PbS QD layer is fully depleted but that the CdS layer is not depleted. However, the measured capacitance deviates from the modeled capacitance bounds and becomes relatively constant for PbS QD thicknesses 97 greater than -150 nm, suggesting that the PbS QD layer is no longer fully depleted. The ~150 nm depletion width implied from the capacitance measurement is consistent with the -170 nm combined diffusion length and depletion width estimated from the optical model. The fact that the highest J,, was obtained with a 160 nm thick layer of PbS QDs also correlates with this characteristic length. As the PbS QD thickness is smaller than this length, the J,, increases with the thickness owing to improved light absorption. However, when the PbS QD thickness exceeds this length, more photons are absorbed in the quasi-neutral region and fewer photons are absorbed in the combined diffusion length and depletion width region, causing inefficient carrier extraction and a lower Jsc. 5.8 CdS Layer Thickness-Dependent Jsc The optical model can be used to guide the device design as well. As mentioned earlier in this chapter, CdS layer can also act as an optical spacer to modulate the optical electric field and concentrate the light absorption in the PbS QD layer. Based on the optical model developed in previous sections, we simulated the Jsc of the PbS QD/CdS devices as a function of CdS layer thicknesses and plotted the results in Figure 5-28. It can be seen that the optimal CdS thickness for maximum Jsc occurs at -50 nm, which is actually the thickness of CdS layer that have been adopted in this study. However, it is also found that the change of the Jsc with respect to CdS layer thicknesses is trivial, which can be rationalized with the distribution of optical electric fields as shown in Figure 5-29. The distances between adjacent Jelectric field12 peaks are roughly equal to the effective carrier extraction range (-170 nm) in the PbS QD layer, suggesting that the concentration of optical electric field and accordingly the variation of Jsc due to the optical spacer (CdS layer) is marginal. 98 16.2 16.0 E E 15.815.6 15.4 15.2 - 15.0 0 14.814.6 0 50 100 150 200 250 300 350 400 CdS Thickness (nm) Figure 5-28. Modeled short-circuit current density of the PbS QD/CdS devices with various CdS layer thicknesses. The thickness of PbS QD layer for this modeling is fixed at 170 nm, which is effective carrier extraction range estimated in the previous sections of this thesis. 99 (a) 301 (b) Al PbS ITO 2 2 401 E 504 C 1.5C 1.5 J+--601 ba C 1 - 701 ~801 0.5 0.5 901 100 lu 200 Distance (nm) 2011 400 3UU Distance (nm) (d) (c) 30, 3 40 1.5E E so 2.5 C 2 ~60 1 00 C 1.5 0 -~70 1 ~80 0 0.5 0.5 90 100 (e) 3C 0 200 100 300 400 0 500 Al CdS ITO PbS 100 200 300 400 500 600 Distance (nm) Distance (nm) 2.5 2 E C 1.5 m 8 1 0 .5 ZOO 400 600 800 Distance (nm) 2 Figure 5-29. Simulated loptical electric field as a function of wavelength and position for various CdS thicknesses: (a) 0 nm, (b) 50 nm, (c) 180 nm, (d) 280 nm, and (e) 400 nm. The thickness of PbS QD layer is fixed at 170 nm. 100 5.9 Conclusion In summary, we demonstrated the fabrication of PbS QD/CdS heterojunction devices based on a low-temperature solution process with good device reproducibility and an average PCE of 3.5%. Thickness-dependent J-V characteristics, optical modeling, and capacitance measurements suggest that the combined diffusion length (~20 nm) and depletion width (150 nm) in the PbS QD layer is ~170 nm. The integration of colloidal QDs and low temperaturedeposited inorganic semiconductors in this work demonstrates the applicability of QDs in a potentially scalable device fabrication process and also opens up the possibility of pairing PbS QDs with other chemical-bath-deposited semiconductors, which may lead to further performance improvement. 5.10 Reference I Brabec, C. J.; Gowrisanker, S.; Halls, J. J. M.; Laird, D.; Jia, S. & Williams, S. P., Polymer-Fullerene Bulk-Heterojunction Solar Cells, Advanced Materials, 2010, 22, 3839-3856 2 Walker, B.; Kim, C. & Nguyen, T.-Q., Small Molecule Solution-Processed Bulk Heterojunction Solar Cells, Chemistry of Materials, 2011, 23, 470-482 3 Kramer, I. J. & Sargent, E. H., Colloidal Quantum Dot Photovoltaics: A Path Forward, ACS Nano, 2011, 5, 8506-8514 4 Boudreault, P.-L. T.; Najari, A. & Leclerc, M., Processable Low-Bandgap Polymers for Photovoltaic Applications, Chemistry of Materials, 2011, 23, 456-469 5 Hines, M. & Scholes, G., Colloidal PbS Nanocrystals with Size-Tunable Near-Infrared Emission: Observation of Post-Synthesis Self-Narrowing of the Particle Size Distribution, Adv. Mater., 2003, 15, 1844-1849 6 Liu, T.-Y.; Li, M.; Ouyang, J.; Zaman, M. B.; Wang, R.; Wu, X.; Yeh, C.-S.; Lin, Q.; Yang, B. & Yu, K., Non-Injection and Low-Temperature Approach to Colloidal Photoluminescent PbS Nanocrystals with Narrow Bandwidth, The Journal of Physical Chemistry C, 2009, 113, 2301-2308 101 7 Moreels, I.; Justo, Y.; De Geyter, B.; Haustraete, K.; Martins, J. C. & Hens, Z., SizeTunable, Bright, and Stable PbS Quantum Dots: A Surface Chemistry Study, ACS Nano, 2011, 5, 2004-2012 8 Shockley, W. & Queisser, H. J., Detailed Balance Limit of Efficiency of p-n Junction Solar Cells, Journal of Applied Physics, 1961, 32, 510-51 9 De Vos, A., Detailed balance limit of the efficiency of tandem solar cells, J. Phys. D: Appl. Phys., 1980, 13, 839-846 10 Johnston, K. W.; Pattantyus-Abraham, A. G.; Clifford, J. P.; Myrskog, S. H.; MacNeil, D. D.; Levina, L. & Sargent, E. H., Schottky-quantum dot photovoltaics for efficient infrared power conversion, Applied Physics Letters, 2008, 92, 151115-151118 11 Ip, A. H.; Thon, S. M.; Hoogland, S.; Voznyy, 0.; 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 Nano, 2012, 7, 577-582 12 Rath, A. K.; Bernechea, M.; Martinez, L.; de Arquer, F. P. G.; Osmond, J. & Konstantatos, G., Solution-processed inorganic bulk nano-heterojunctions and their application to solar cells, Nat Photon, 2012, 6, 529-534 13 Tang, J.; Liu, H.; Zhitomirsky, D.; Hoogland, S.; Wang, X.; Furukawa, M.; Levina, L. & Sargent, E. H., Quantum Junction Solar Cells, Nano Letters, 2012, 12, 4889-4894 14 Oladeji, I. 0. & Chow, L., Optimization of Chemical Bath Deposited Cadmium Sulfide Thin Films, Journal of The Electrochemical Society, ECS, 1997, 144, 2342-2346 15 Mane, R. & Lokhande, C., Chemical deposition method for metal chalcogenide thin films, Materials Chemistry and Physics, 2000, 65, 1 - 31 16 Britt, J. & Ferekides, C., Thin-film CdS/CdTe solar cell with 15.8% efficiency, Applied Physics Letters, 1993, 62, 2851-2852 17 Contreras, M. A.; Romero, M. J.; To, B.; Hasoon, F.; Noufi, R.; Ward, S. & Ramanathan, K., Optimization of CBD CdS process in high-efficiency Cu(In,Ga)Se2-based solar cells, Thin Solid Films, 2002, 403-404, 204-211 18 Bber, K. W., Cadmium sulfide enhances solar cell efficiency, Energy Conversion and Management, 2011, 52, 426 - 430 102 19 Pattantyus-Abraham, A. G.; Kramer, I. J.; Barkhouse, A. R.; Wang, X.; Konstantatos, G.; Debnath, R.; Levina, L.; Raabe, I.; Nazeeruddin, M. K.; Gratzel, M. & Sargent, E. H., Depleted-Heterojunction Colloidal Quantum Dot Solar Cells, ACS Nano, 2010, 4, 3374- 3380 20 Hyun, B.-R.; Zhong, Y.-W.; Bartnik, A. C.; Sun, L.; Abrufna, H. D.; Wise, F. W.; Goodreau, J. D.; Matthews, J. R.; Leslie, T. M. & Borrelli, N. F., Electron Injection from Colloidal PbS Quantum Dots into Titanium Dioxide Nanoparticles, ACS Nano, 2008, 2, 2206-2212 21 Mumford, P. D. & Cahay, M., Space-charge effects and current self-quenching in a metal/CdS/LaS cold cathode, Journal of Applied Physics, 1998, 84, 2754-2767 22 Luther, J. M.; Law, M.; Beard, M. C.; Song, Q.; Reese, M. 0.; Ellingson, R. J. & Nozik, A. J., Schottky Solar Cells Based on Colloidal Nanocrystal Films, Nano Letters, 2008, 8, 3488-3492 23 Hu, M.; Noda, S.; Okubo, T.; Yamaguchi, Y. & Komiyama, H., Structure and morphology of self-assembled 3-mercaptopropyltrimethoxysilane layers on silicon oxide, Applied Surface Science, 2001, 181, 307 - 316 24 Morris-Cohen, A. J.; Donakowski, M. D.; Knowles, K. E. & Weiss, E. A., The Effect of a Common Purification Procedure on the Chemical Composition of the Surfaces of CdSe Quantum Dots Synthesized with Trioctylphosphine Oxide, The Journal of Physical Chemistry C, 2010, 114, 897-906 25 Pattantyus-Abraham, A. G.; Kramer, I. J.; Barkhouse, A. R.; Wang, X.; Konstantatos, G.; Debnath, R.; Levina, L.; Raabe, I.; Nazeeruddin, M. K.; Gratzel, M. & Sargent, E. H., Depleted-Heterojunction Colloidal Quantum Dot Solar Cells, ACS Nano, 2010, 4, 33743380 26 Gao, J.; Luther, J. M.; Semonin, 0. E.; Ellingson, R. J.; Nozik, A. J. & Beard, M. C., Quantum Dot Size Dependent J-V Characteristics in Heterojunction ZnO/PbS Quantum Dot Solar Cells, Nano Letters, 2011, 11, 1002-100 27 Brown, P. R.; Lunt, R. R.; Zhao, N.; Osedach, T. P.; Wanger, D. D.; Chang, L.-Y.; Bawendi, M. G. & Bulovi6, V., Improved Current Extraction from ZnO/PbS Quantum Dot Heterojunction Photovoltaics Using a MoO3 Interfacial Layer, Nano Letters, 2011, 11, 2955-2961 103 28 Sze, S.M. & Ng, K.K., Physcis of Semidconductor Devices, John Wiley & Sons, 2010, 719-736 29 Peumans, P.; Yakimov, A. & Forrest, S. R., Small molecular weight organic thin-film photodetectors and solar cells, Journal of Applied Physics, 2003, 93, 3693-3723 30 Law, M.; Beard, M. C.; Choi, S.; Luther, J. M.; Hanna, M. C. & Nozik, A. J., Determining the Internal Quantum Efficiency of PbSe Nanocrystal Solar Cells with the Aid of an Optical Model, Nano Letters, 2008, 8, 3904-3910 31 Kim, J.; Kim, S.; Lee, H.-H.; Lee, K.; Ma, W.; Gong, X. & Heeger, A., New Architecture for High-Efficiency Polymer Photovoltaic Cells Using Solution-Based Titanium Oxide as an Optical Spacer, Advanced Materials, 2006, 18, 572-576 32 Pettersson, L. A. A.; Roman, L. S. & Inganas, 0., Modeling photocurrent action spectra of photovoltaic devices based on organic thin films, Journal of Applied Physics, 1999, 86, 487-496 33 Burkhard, G. F.; Hoke, E. T. & McGehee, M. D., Accounting for Interference, Scattering, and Electrode Absorption to Make Accurate Internal Quantum Efficiency Measurements in Organic and Other Thin Solar Cells, Advanced Materials, 2010, 22, 3293-3297 34 Tang, J.; Brzozowski, L.; Barkhouse, D. A. R.; Wang, X.; Debnath, R.; Wolowiec, R.; Palmiano, E.; Levina, L.; Pattantyus-Abraham, A. G.; Jamakosmanovic, D. & Sargent, E. H., Quantum Dot Photovoltaics in the Extreme Quantum Confinement Regime: The Surface-Chemical Origins of Exceptional Air- and Light-Stability, ACS Nano, 2010, 4, 869-878 35 Rath, A. K.; Bernechea, M.; Martinez, L. & Konstantatos, G., Solution-Processed Heterojunction Solar Cells Based on p-type PbS Quantum Dots and n-type Bi2S3 Nanocrystals, Advanced Materials, 2011, 23, 3712-3717 36 Green, M. A., Solar Cells, Englewood Cliffs, NJ : Prentice-Hall, 1998, 67-68 37 Dirote, E. V., Nanotechnology at the Leading Edge, 2006, 109 104 Chapter 6 Efficiency Improvement in Inorganic-Organic Hybrid Photovoltaic Devices via Solvent-Assisted Self Assembly 6.1 Introduction Solution-processed nanostructured hybrid solar cells consisting of organic and inorganic components have recently attracted considerable interest due to the combined advantages of both material classes [1, 2]. In hybrid organic-inorganic bulk heterojunction (BHJ) solar cells [3, 4], electron-donating conijugated polymers are blended with inorganic semiconductor nanomaterials, taking advantage of their solution processability [5, 6]. So far, a variety of nanomaterials such as GaAs [7], CdSe [8], PbS [9], TiO 2 , and ZnO [10, 11], nanowires or nanocrystals have been investigated for applications in hybrid solar cells. To improve BHJ solar cell device efficiency, several design criteria have to be taken into consideration. The domain size of the organic donor and inorganic acceptor materials should be comparable or smaller than the exciton diffusion length to increase the probability of exciton dissociation across the heterojunction interface. For many conjugated polymers, the exciton diffusion length is less than 10 nm [12], but in standard BHJ nanostructured solar cells processed from solution, device morphology cannot be precisely controlled on such a small scale. In addition, efficient transport of charge carriers to their respective electrodes before recombination is the key for high device current and voltage. Therefore, high electron and hole mobilities, controllable nanomorphology, and a well-structured interface are critical factors for developing efficient hybrid solar cells [13]. QDs are particularly well-suited for applications in hybrid BHJ solar cells because their absorption can be tuned to cover a broad spectral range. QDs can have relatively high electron mobility [14], and good photo and chemical stability. However, achieving a controllable bicontinuous percolation network and a well-controlled interface between QDs and the polymer matrix remain challenging due to the chemical difference between inorganic and organic components. In addition, the inorganic QDs tend to vertically phase separate from non-polar conjugated polymers at higher loading concentrations, decreasing the interfacial area [15], and this phase separation of the photoactive films can be only partly mitigated by selection of solvent 105 and spin-coating conditions. Therefore, the next generation of hybrid QD photovoltaics will require a strategy for controlling phase separation, increasing the interfacial area, and improving the optoelectronic interactions of inorganic QDs and organic polymers. Here, we demonstrate a facile method to synthesize and control the nanoscale morphology of P3HT nanofiber and CdS QD hybrids using solvent-assisted self-assembly [16] and ligand exchange methods [17]. Solvent-assisted self assembly and ligand exchange were used to control the interface of P3HT nanofiber/CdS QD nano-hybrids and the CdS QD interparticle distance, respectively. The effects of controlling morphology on the optical and electrical properties are discussed. We find that the formation of an interpenetrating and percolating P3HT nanofiber/CdS QD BHJ network gives rise to efficient charge transfer and charge transport, significantly improving photovoltaic performance. 6.2 Device Structure and Fabrication The synthesis of CdS QDs followed procedures reported by Peng et al. [18] and is described in Chapter 3. After the synthesis, the CdS QDs were centrifuged, precipitated with isopropanol, and redispersed in hexane. P3HT nanofibers were prepared using solution-phase self assembly. Nonpolar P3HT was initially completely solvated in a nonpolar solvent, 1,2dichlorobenzene (1,2-DCB). To weaken the interaction between P3HT and the solvent, the solvent polarity was increased by adding a marginal solvent, cyclohexanone. As a consequence, the interaction among P3HT polymer chains becomes dominant, puffing P3HT into an aggregated state. In this stage, the P3HT solution forms a dark purple gel whose viscosity can be strongly decreased under manual agitation, indicating the formation of P3HT nanofibers. To form the photoactive layer of the hybrid solar cell whose structure is shown in Figure 6-1(a), P3HT nanofibers were blended with CdS QDs. We selected CdS QDs due to the wide offset between the HOMO level of P3HT and the conduction band edge of CdS QDs, which can potentially provide large Voc (Figure 6-1(b)) [19]. Hybrid solar cells consisting of P3HT nanofiber/CdS QD random-mixed and self-assembled structures were prepared by a solventassisted method. For the random-mix process, the same solvent, 1,2-DCB, was used for CdS QDs and P3HT nanofibers. On the contrast, in the self-assembly process, the 1-D coaxial nanofiber/QD structures were formed by first dissolving the P3HT nanofibers and CdS QDs into 1,2-DCB and octane, respectively, and then mixing these two solutions together. The P3HT 106 nanofiber/CdS QD blended solution was spun onto ITO-coated glass substrates to form a 140-nm thick film. The coated devices were heat-treated at 175 *C for 10 min and quickly cooled to room temperature. Finally, the 10-nm thick bathocuproine (BCP) hole blocking layer and the top Mg/Ag electrode were evaporated and the final device area was defined as the overlap between the top and bottom electrodes. (a) (b) Mg:Ag ITO - - - -PEDO 4.8 eV 5.0 eV ITO Glass 3.7 eV P3HT CdS BCP 6.6eV Figure 6-1. (a) The schematic device architecture and (b) the corresponding flat-band energy diagram. 6.3 Material Characterization: Random Mix versus Self Assembly The photoactive material of our hybrid solar cells consists of the electron-donor P3HT nanofibers and electron-acceptor CdS QDs. Figure 6-2(a), (b) shows representative TEM images of P3HT nanofibers and colloidal CdS QDs (-4 nm in diameter) capped with oleic acid ligands, respectively. As prepared P3HT nanofibers have an aspect ratio of 100 with an average length of 1000±150 nm and diameter of 10±2 nm. Figure 6-3 shows the UV-VIS absorbance spectra of P3HT and CdS QD components. In comparison with amorphous P3HT phase, a significant redshift and an additional vibronic feature (600 nm) are observed in the absorbance spectrum of P3HT nanofibers (Figure 6-3(a)). These results indicate the formation of I-D self-organized P3HT nanofibers with improved structural ordering. It has been reported that the n-n interactions between polymer chains lead to superior electronic properties with a two-order-of-magnitude increase of the hole mobility in P3HT nanofibers [20]. 107 (a) (b) Figure 6-2. TEM images of (a) P3HT nanofibers and (b) CdS QDs. 1.0 (b) (a) 1.0 - nanofiber amorphous 0.8- 0.8- 0.6 - ( 0.6 C .2. 2 ~~0. CU 0 0 0.2-6 ~0.2 0.0 300 400 5 6 700 Wavelength (nm) 800 0.0 350 400 450 500 550 600 Wavelength (nm) Figure 6-3. UV-VIS absorbance spectra of (a) amorphous P3HT, P3HT nanofribers, and (b) CdS QDs. Because the diameters of P3HT nanofibers are below the exciton diffusion length of P3HT phase, we studied the effect of the self-assembly process on exciton dissociation between P3HT nanofibers and CdS QDs by comparing a randomly blended system (random-mix process) with 1-D coaxial nanostructures formed by decorating P3HT nanofibers with CdS QDs (selfassembly process). A key to the self-assembly process is to introduce a poor solvent, allowing P3HT nanofibers and CdS QDs to form aggregates to minimize their surface energy. In the present study, the P3HT nanofibers and CdS QDs were dissolved in 1,2-DCB and in octane, respectively. Octane is known to be a good solvent for CdS QDs but a poor solvent for P3HT. 108 Therefore, we expect CdS QDs preferentially attach to the surface of P3HT nanofibers to reduce surface energy. Figure 6-4(a) shows a TEM image of the randomly blended P3HT/CdS system; the CdS QDs are randomly distributed and there is no obvious ordering interaction between P3HT nanofibers and CdS QDs. To increase the organic-inorganic interaction and maximize the interfacial area, we self-assembled CdS QDs onto one-dimensional (1-D) P3HT nanofibers, producing P3HT nanofiber/CdS QD coaxial hybrids (Figure 6-4(b)). This architecture carries advantages over a randomly distributed system because phase segregation is minimized and the P3HT nanofiber/QD interface is maximized, which can potentially improve the dissociation of photo-generated excitons in P3HT nanofibers due to the conduction band offset at the interface. Figure 6-5 shows the film absorbance and solution photoluminescence (PL) spectra produced from random-mixed and self-assembled P3HT nanofibers with 80 wt% CdS QDs. The vibronic absorption feature at ~600 nm in the self-assembled film is clearly resolved, indicating that the crystalline P3HT nanofibers are preserved after CdS QD self assembly. The PL intensity of the P3HT nanofiber/CdS QD self-assembled film is significantly quenched, presumably due to the efficient nonradiative channel for charge transfer from P3HT nanofibers to CdS QDs. (a) (b)-'1 ' Figure 6-4. TEM images of P3HT nanofiber/CdS QD film synthesized (a) without self-assembly (random-mix process) and (b) using a solvent-assisted self-assembly process. The inset images show schematic representations of each; the random-mix and self-assembly method is used to control the interface between CdS QDs (yellow spheres) and P3HT nanofibers (purple lines). 109 (a) (b) 0.8 7 Random m ix 6 0. 0.6 0 e (au) C C 3.2 5 4 32- 400 300 350 400 450 500 550 600 650 700 Wavelength (nm) j Self-assem bly Soo 600 700 800 900 Wavelength (nm) Figure 6-5. (a) An UV-VIS absorbance spectrum of self-assembled P3HT nanofiber/CdS QD film. (b) Solution PL spectra of the random-mixed and self-assembled P3HT nanofiber/CdS QD samples. In order to understand the bonding structure of the self-assembled P3HT nanofiber/CdS QD hybrids, XPS spectra of the samples without and with self-assembly were obtained (Figure 6-6(a,b), respectively). We focus on the XPS spectra of S 2p peaks to more precisely identify the bond formed during self assembly. Figure 6-6(a) shows the S 2p peaks for random-mixed samples where the S bonding energies is well matched through peak deconvolution with the independent P3HT and CdS components. In comparison, XPS spectrum of the self-assembled P3HT nanofiber/CdS QD hybrids (Figure 6-6(b)) reveals a new intermediate bonding energy which can be assigned to the C-S-Cd bond, connecting the P3HT polymer backbone and CdS QDs [21]. These findings are further supported by transient PL measurements of P3HT nanofiber/CdS QD hybrids with and without self-assembly (Figure 6-7), which show that selfassembled samples have a shorter exciton lifetime and thus faster exciton dissociation. Taken together, these results can be understood as follows: the self assembly of CdS QDs onto the surface of highly ordered P3HT nanofibers increases the interaction between these two materials and creates a large number of interfaces for charge transfer [22], which results in a shorter PL lifetime, quenched PL intensities, and increased charge transfer rates. 110 Random-mixed (b) Self- CsCd o 5. assembled Mo4 3P3HT 3 A C 162 A 2 - 162 156 158 160 Binding Energy (eV) 156 160 158 Binding Energy (eV) Figure 6-6. XPS spectra of high resolution S 2p in (a) random-mixed and (b) self-assembled P3HT nanofiber/CdS QD film. 10 -- Self-assembled Random-mixed 10 10 10n 4-J 10 10 U -D N 1024 - E 0 104 -10 0 10 20 30 40 50 Time (ns) Figure 6-7. Time-resolved photoluminescence spectra of the random-mixed and self-assembled P3HT nanofiber/CdS QD film. The film was excited by a pulsed diode laser (PicoQuant) at 414 nm, 2.5 MfHz. Emission from the film was collected by an avalanche photodiode (PicoQuant) with a 460nm band-pass filter. Photoluminescence decay of the sample was measured with a time-correlated single photon counting module (PicoQuant). It was recently demonstrated that the performance of solar cells based on PbS QDs can be improved by ligand-exchange treatments involving thiols [23]. The beneficial effect of thiols has 111 been attributed to a shortened QD spacing, improved cross-linking of QDs, and improved surface passivation. The photoconductivity of different QDs has been shown to be largely enhanced with a consequence of the ligand exchange treatment [17]. We performed the ligand exchange in our organic P3HT nanofiber-inorganic CdS QD hybrid system to shorten the interdistance between CdS QDs and P3HT nanofibers and among CdS QDs themselves. The CdS QDs were originally capped with oleic acid and n-butylamine surfactants, which were removed through an ethanedithiol (EDT) treatment in acetonitrile solvent. Acetonitrile is a relatively good solvent for ligands, but a poor solvent for P3HT nanofibers. The solvent quality induces ordering in the polymer phase and changes the degree of phase separation. The surface morphology before and after the ligand exchange was investigated using AFM and TEM (Figure 6-8(a)-(c), respectively) and it was found that the P3HT nanofiber/CdS QD film forms small agglomerates after the ligand exchange, which indicates that CdS QDs get closer to one another. Figure 6-8. Tapping mode AFM images of P3HT nanofiber/CdS QD thin film (a) before and (b) after EDT ligand exchange. (c) TEM image of curved P3HT NWs after the ligand exchange. 6.4 Device Characterization Next, we examined the effect of the controlled nanoscale morphology on electrical properties. To directly measure charge recombination rates we fabricated a device after self assembly but with and without ligand-exchange treatment, and performed small perturbation transient V,, decay measurements [24, 25]. In this measurement, a low-intensity pulsed laser light at 520 nm wavelength is used to induce a small perturbation to the Vc by transiently generating additional electrons and holes. The resulting additional transient photovoltage (V) then decays with a lifetime that is determined by the recombination rate constant (kre) of the electrons and holes. In Figure 6-9, kec is shown as a function of the solar simulator light intensity. 112 For all devices, we observe that the EDT-treated device exhibits k,,c approximately two times higher than that of the non-treated device, which is likely due to the fact that photogenerated charge carrier concentrations of EDT-treated devices are significantly higher than that of the pristine device. The ligand exchange process reduces the distance among CdS QDs and P3HT nanofibers and thus increases the concentration of free charge carriers due to more efficient exciton dissociation, consequently increasing the charge recombination rate constant. The charge carrier concentration and charge transport is highly dependent on the carrier mobility of the materials, which also strongly influences the photovoltaic performance. Therefore, both electron and hole mobility in CdS and P3HT components should be matched and high enough to extract charges efficiently and prevent build-up of space charge. We have tested the electron-only and hole-only devices with self-assembled hybrid film (Figure 6-10), which indicates the wellmatched electron and hole mobility after the EDT treatment process. We note that the conductivity of P3HT nanofibers is reduced by the EDT treatment, which may result from the decreased P3HT molecular ordering as can be manifested as the curving features shown in the TEM image (Figure 6-8(c)). 6x1 03 5x 10 4x103 3x103 S2x1 0 Pristine EQ 103 10 100 Power Dens ity (mnW/cm) Figure 6-9. Charge carrier recombination rate constants krec of pristine and EDT-treated devices determined by transient open circuit voltage decay measurements at different illumination levels. 113 (a) 0.004 (b) 0.'014 electron W EDT electron w/b EDT 0.003 0. 03 0.002 0. 02 E 0.001 E 0. 01 0.000 0. 00' -0.001 -0. 01 0 -0.002 -hoe/EDT hole w/o E DT - 0 -0. 02 -0.003 -0. 03- -_n flA 5 -4 -3 -2 -1 0 1 Voltage (V) 2 3 4 5 -4 -3 -2 -1 0 1 Voltage (V) 2 3 4 5 Figure 6-10. (a) J-V characteristics of an electron-only P3HT nanofiber/CdS QD device before and after EDT treatment. (b) J-V characteristics of a hole-only P3HT nanofiber/CdS QD device before and after EDT treatment. The effect of EDT treatment and P3HT nanofiber curving on the transport properties of P3HT nanofiber/CdS QD film was demonstrated by the DC J-V characteristics of the electron-only and hole-only devices. The electron-only device consisted of Ag:self-assembled monolayer (SAM), P3HT nanofiber/CdS QD, and Mg:Ag layers. The Ag:SAM layer blocks the hole injection [26]. The treated electron-only device shows a strong enhancement of electron transport over the untreated device. The hole-only device consisted of ITO, PEDOT:PSS, P3HT nanofiber/CdS QD, and Au layers, where the Au layer prevents electron injection. The treated device showed a decrease in current density. This result suggests that the EDT treatment reduces the hole transport. Following these promising measurements, devices based on self-assembled and EDTtreated films were prepared with the structure described in Figure 6-1(a). The bright-field TEM image of a cross-section of a typical complete device (Figure 6-11(a)) shows that CdS QDs were uniformly distributed across the 150 nm thick photoactive layer. The self assembly of CdS QDs onto P3HT nanofibers is confirmed using energy dispersive x-ray spectroscopy elemental mapping image (Figure 6-11(b) and Figure 6-12). In comparison with the QD vertical phase separation that often occurs in randomly blended inorganic-organic hybrid devices [15], the image suggests that the P3HT nanofiber structure and self-assembly process plays a crucial role in templating CdS QDs across the vertical direction and improving the interface between P3HT nanofibers and CdS QDs. 114 CP Cd$.-P3HT Figure 6-11. (a) Cross-sectional TEM image of P3HT nanofiber/CdS QD hybrid solar cells. (b) Energy dispersive X-ray spectroscopy elemental mapping (Red color: Cd; Green color: In; and Blue: Ag) of the image shown in (a). Figure 6-12. Cross-sectional TEM EDS mapping image of the ITO/PEDOT:PSS/P3HT:CdS/BCP/Mg:Ag hybrid solar cell device. (a)-(d): The Mg, Cd, S, and In element mapping images. Figure 6-13 shows the current density-voltage characteristics of the P3HT nanofiber/CdS QD hybrid photovoltaic devices using random-mixed film and self-assembled film as a photoactive layer. The device at 80 wt% CdS QDs without self assembly shows Jsc of 1.92 115 mA/cm 2 , Voc of 0.78 V, and a FF of 0.41 under simulated AM 1.5 illumination (100 mW/cm 2 ), and gives power conversion efficiency of 0.61%. In contrast, the device with the self-assembly process exhibits higher Jsc of 5.75 mA/cm 2, Voc of 1.00 V, a FF of 0.39, and power conversion efficiency of 2.24%. The enhancement of the photovoltaic performance in the device with the self-assembly process was attributed to the increased charge transfer efficiency in the photoactive film, enabled by self-assembling CdS QDs onto highly ordered P3HT nanofibers as evidenced in previous sections. 2 Eo E> n -2 4- 40 -6 Random-mixed Self-assembled -8 -10 -- - - - - -I -0.4 -0.2 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 Voltage (V) Figure 6-13. J-V characteristics of random-mixed (blue) and self-assembled (red) P3HT nanofiber/CdS QD solar cells. Figure 6-14 (a) shows the current density-voltage characteristics of the pristine and EDTtreated P3HT nanofiber/CdS QD hybrid photovoltaic devices, and Figure 6-14(b) shows a summary of the Jsc and Voc as a function of CdS wt% for the samples after the self assembly and EDT treatment. Our results show that a significant improvement of the photovoltaic performance is achieved in hybrid devices fabricated using self assembly and EDT treatment processes, giving an improved Jsc of 10.96 mA/cm 2, Voc of 1.06 V, FF of 0.35 and the maximum efficiency of 4.06% (average power conversion efficiency of 3.2% measured over 100 devices). The improvement of the performance in the device after EDT treatment originated from the enhanced charge transfer efficiency and charge collection efficiency provided by shorter 116 P3HT nanofiber/CdS QD distances and QD/QD distances, which has been supported by the microscopy images (Figure 6-8) and conductivity measurement (Figure 6-10) presented in previous sections. (a) 4 . , . , . , , , . (b) 12 . 2 E E 4-J LI (U 1+~*Voc 0 -2 -4t -6 -8- 0.7 8 04 -10-12 - Pristine -14 - EDT-treated -16 - ' - ' - ' - ' - ' . ' -0.4 -0.2 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.1 -T 1.0 0.9 0.8E 2. 0.6> 0.5 0.4 1.4 Voltage (V) 50 60 70 80 90 CdS Weight Concentration (W/o) Figure 6-14. (a) J-V characteristics of pristine (red) and EDT-treated (green) P3HT nanofiber/CdS QD solar cells. (b) The photovoltaic performance summary (Jsc and Voc) of the EDT-treated and self-assembled devices as a function of the CdS weight concentration. 6.5 Conclusion In summary, we have presented a facile and simple method to synthesize and control the nanomorphology of P3HT/CdS coaxial nanofiber BHJ hybrid solar cells. The self assembly of CdS QDs onto crystalline P3HT nanofibers results in not only the controllable dispersion of CdS QDs within the P3HT matrix, but also stronger electronic interaction between CdS QDs and P3HT nanofibers. The formation of bicontinuous donor/acceptor phases and a well-defined interface in the hybrid photoactive film through self assembly and ligand exchange can largely enhance charge transfer and transport efficiency, which are essential for hybrid inorganic-organic solar cells. The results shown here represent a direction for increasing the efficiency of hybrid solar cells via enhancing the interfacial interactions and controlling nanomorphology between the organic and inorganic materials. 117 Author Contributions This chapter is an equal contribution between Dr. Shenqiang Ren and' Liang-Yi Chang. Dr. Shengqiang Ren is responsible for the device fabrication and electrical measurements. LiangYi Chang is responsible for the material synthesis and material characterization of the work. 6.6 Reference 1 Milliron, D. J.; Gur, I. & Alivisatos, A. P., Hybrid Organic-Nanocrystal Solar Cells, MRS Bulletin, 2005, 30, 41-44 2 Gur, I.; Fromer, N. A.; Chen, C.-P.; Kanaras, A. G. & Alivisatos, A. P., Hybrid Solar Cells with Prescribed Nanoscale Morphologies Based on Hyperbranched Semiconductor Nanocrystals, Nano Letters, 2007, 7, 409-414 3 Yu, G. & Heeger, A. J., Charge separation and photovoltaic conversion in polymer composites with internal donor/acceptor heterojunctions, Journal of Applied Physics, 1995, 78, 4510-4515 4 Huynh, W. U.; Dittmer, J. J.; Alivisatos, A. P., Hybrid nanorod-Polymer Solar Cells, Science, 2002, 295, 2425-2427 5 Greenham, N. C.; Peng, X. G.; Alivisatos, A. P., Charge separation and transport in conjugated-polymer/semiconductor-nanocrystal composites studied by photoluminescence quenching and photoconductivity, Phys. Rev. B, 1996, 54, 1762817637 6 Zhou, Y.; Riehle, F. S.; Yuan, Y.; Schleiermacher, H.-F.; Niggemann, M.; Urban, G. A. & Kruger, M., Improved efficiency of hybrid solar cells based on non-ligand-exchanged CdSe quantum dots and poly(3-hexylthiophene), Applied Physics Letters, 2010, 96, 013304 7 Ren, S.; Zhao, N.; Crawford, S. C.; Tambe, M.; Bulovic, V. & Gradecak, S. Heterojunction Photovoltaics Using GaAs Nanowires and Conjugated Polymers Nano Letters, 2011, 11, 408-413 8 Wang, P.; Abrusci, A.; Wong, H. M. P.; Svensson, M.; Andersson, M. R. & Greenham, N. C., Photoinduced Charge Transfer and Efficient Solar Energy Conversion in a Blend of a Red Polyfluorene Copolymer with CdSe Nanoparticles, Nano Letters, 2006, 6, 17891793 118 9 Zhao, N.; Osedach, T. P.; Chang, L.-Y.; Geyer, S. M.; Wanger, D.; Binda, M. T.; Arango, A. C.; Bawendi, M. G. & Bulovic, V., Colloidal PbS Quantum Dot Solar Cells with High Fill Factor, ACS Nano, 2010, 4, 3743-3752 10 Shim, M. & Guyot-Sionnest, P., Organic-Capped ZnO Nanocrystals: Synthesis and nType Character, Journal of the American Chemical Society, 2001, 123, 11651-11654 11 Beek, W.; Wienk, M. & Janssen, R., Hybrid Solar Cells from Regioregular Polythiophene and ZnO Nanoparticles, Advanced Functional Materials, 2006, 16, 1112- 1116 12 Shaw, P. E.; Ruseckas, A. & Samuel, I. D. W., Exciton Diffusion Measurements in Poly(3-hexylthiophene), Advanced Materials, 2008, 20, 3516-3520 13 Chen, L.-M.; Hong, Z.; Li, G. & Yang, Y., Recent Progress in Polymer Solar Cells: Manipulation of Polymer:Fullerene Morphology and the Formation of Efficient Inverted Polymer Solar Cells, Advanced Materials, 2009, 21, 1434-1449 14 Kovalenko, M. V.; Scheele, M.; Talapin, D. V., Colloidal Nanocrystals with Molecular Metal Chalcogenide Surface Ligands, Science, 2009, 324, 1417-1420 15 Coe-Sullivan, S.; Steckel, J.; Woo, W.-K.; Bawendi, M. & Bulovic, V., Large-Area Ordered Quantum-Dot Monolayers via Phase Separation During Spin-Casting, Advanced Functional Materials, 2005, 15, 1117-1124 16 Yang, S.; Wang, C.-F. & Chen, S., Interface-Directed Assembly of One-Dimensional Ordered Architecture from Quantum Dots Guest and Polymer Host, Journal of the American Chemical Society, 2011, 133, 8412-8415 17 Jarosz, M. V.; Porter, V. J.; Fisher, B. R.; Kastner, M. A.; Bawendi, M. G., Photoconductivity studies of treated CdSe quantum dot films exhibiting increased exciton ionization efficiency, Phys. Rev. B, 2004, 70, 195327-195339 18 Peng, Z. A. & Peng, X., Formation of High-Quality CdTe, CdSe, and CdS Nanocrystals Using CdO as Precursor, Journal of the American Chemical Society, 2001, 123, 183-184 19 Leventis, H. C.; King, S. P.; Sudlow, A.; Hill, M. S.; Molloy, K. C. & Haque, S. A., Nanostructured Hybrid Polymer/Inorganic Solar Cell Active Layers Formed by Controllable in Situ Growth of Semiconducting Sulfide Networks, Nano Letters, 2010, 10, 1253-1258 119 20 Park, Y. D.; Lee, S. G.; Lee, H. S.; Kwak, D.; Lee, D. H.; Cho, K., Solubility-driven polythiophene nanowires and their electrical characteristics, J. Mater. Chem., 2010, 21, 2338-2343 21 NIST X-ray Photoelectron Spectroscopy Database, V. N. I. o. S. a. T., Gaithersburg, 2003; http://srdata.nist.gov/xps 22 Zhao, L.; Pang, X.; Adhikary, R.; Petrich, J. W. & Lin, Z., Semiconductor Anisotropic Nanocomposites Obtained by Directly Coupling Conjugated Polymers with Quantum Rods, Angewandte Chemie International Edition, 2011, 50, 3958-3962 23 Luther, J. M.; Law, M.; Song, Q.; Perkins, C. L.; Beard, M. C. & Nozik, A. J., Structural, Optical, and Electrical Properties of Self-Assembled Films of PbSe Nanocrystals Treated with 1,2-Ethanedithiol, ACS Nano, 2008, 2, 271-280 24 O'Regan, B. C.; Scully, S.; Mayer, A. C.; Palomares, E. & Durrant, J., The Effect of A1203 Barrier Layers in TiO2/Dye/CuSCN Photovoltaic Cells Explored by Recombination and DOS Characterization Using Transient Photovoltage Measurements, The Journal of Physical Chemistry B, 2005, 109, 4616-4623 25 Goh, C.; Scully, S. R. & McGehee, M. D., Effects of molecular interface modification in hybrid organic-inorganic photovoltaic cells, Journal of Applied Physics, 2007, 101, 114503 26 de Boer, B.; Hadipour, A.; Mandoc, M.; van Woudenbergh, T. & Blom, P., Tuning of Metal Work Functions with Self-Assembled Monolayers, Advanced Materials, 2005, 17, 621-625 120 Chapter 7 Outlook 7.1 Conclusion Colloidal QDs and organic semiconductors are promising building blocks for lowtemperature, solution-processed, and hence low-cost solar cells. This work developed the solar cell employing these two material systems with the focus on the junction interface-pn heterojunction in an all-inorganic solar cell and donor/acceptor interface in an organic/inorganic solar cell. The findings and contributions of this thesis are summarized below: a) Dip coating provides the best film quality (in terms of surface morphology) and is preferred in the application where a defect-free film is a must, such as PbS QD/CdS bilayer heterojunction devices. However, in most cases, such as QD Schottky devices, ZnO/PbS QD devices, and inorganic/organic hybrid QD devices, spin coating is good enough and results in device performance on a bar with that of dip-coated devices. On the other hand, the film deposited by spray coating is considerably rougher and can cause leakage or shunting issues in thin devices, but we also noticed that spray coating is an economic deposition technique in view of its solution/material consumption. Therefore, spray coating may be an attractive technique when large-area and thick coating is essential. b) Prior to this work, PbS or PbSe QD heterojunction photovoltaic devices relied on sintered or sputtered n-type transition metal oxides such as ZnO and TiO 2 to form rectifying junctions with p-type PbS or PbSe QDs. This thesis demonstrated a new junction, namely PbS QD/CdS, with a low-temperature solution process that has not been employed in this field before. The resulted PCE was comparable to that of devices based on sputtered or sintered ZnO when using similar dithiol treatments on PbS or PbSe QDs. c) The measured thickness-dependent Jsc and device optical modeling suggested a combined exciton/carrier diffusion length and depletion width of -170 nm, consistent with the depletion width (~150 nm) measured by capacitance measurement, and implied a short exciton/carrier diffusion length of~20 nm. 121 d) The PbS QD/CdS device yield and reproducibility issue stemming from pinholes was successfully addressed, and a procedure for the fabrication of highly-reproducible devices was created. e) P3HT nanofibers were pre-formed in a good solvent (1,2-DCB) by the addition of a marginal solvent (cyclohexanone). The resultant thin film spin-cast from the nanofiber solution created a three-dimensional percolation network for carrier collection. f) P3HT nanofiber/CdS QD blended film prepared with mixed solvents (octane and 1,2-DCB) revealed preferential decoration of CdS QDs on P3HT nanofibers and stronger interaction between CdS QDs and P3HT nanofibers as opposed to the film prepared with single solvent (1,2-DCB). The J-V characteristic of the device fabricated with the mixed-solvent scheme also exhibited better photovoltaic performance than that of the device fabricated with single solvent. 7.2 Future Work Although reasonable PCE has been demonstrated using a low-temperature solution process, further investigations are required for closing the gap between the current performance and the threshold performance for commercialization. Possible improvements can be achieved through both the device structure and material aspects which are outlined below: a) Carrier concentration in PbS QD layer was shown to increase by five-fold after exposing to the aqueous bath during CdS deposition, leading to a narrower depletion width and less efficient carrier separation in the layer. Therefore, to avoid exposing PbS QDs to the aqueous bath, the development of an inverted CdS/PbS QD device in which the CdS layer will be deposited prior to the PbS QD layer is proposed. b) As mentioned in Chapter 2, chemical treatment can modulate doping concentration, carrier mobility, and trap states in QDs. Consequently, a treatment that can realize graded doping to increase built-in potential and simultaneously maintain low doping level in the absorber layer, that can improve carrier mobility, and that can diminish the density of trap states, especially those deep in the band gap, should be optimized and integrated into the device processing steps. c) On the other hand, carrier concentration in the window layer (CdS) should be raised to increase built-in potential and extend the depletion region in the absorber layer. Other 122 chemical-bath-deposited semiconductors can also be explored to form rectifying junctions with PbS QDs. d) The crossover between photo and dark J-V characteristics may originate from the previouslyreported Schottky barrier between ITO and PbS QD layers, opposing to the main diode between PbS QD and CdS layers in the PbS QD/CdS device or between PbS QD and Al layers in the PbS QD Schottky device. Therefore, the interface between ITO and PbS QDs should be further engineered (e.g. insertion of a buffer layer, MoO 3 ) to reduce contact barriers. e) In the P3HT nanofiber/CdS QD hybrid device, the morphology of CdS QDs can be experimented (such as the employment of nanorods) to improve carrier transport through the QD phase. f) Lower-band-gap acceptor QDs can also replace CdS QDs to increase light absorption in the red and near-infrared regions. g) Once the efficiency of the QD photovoltaic device approaches the threshold (~10 % under favorite policies and economics) for commercialization, the lifetime of the device has to be shown to be at least 10 years, therefore requiring substantial efforts to achieve the stability under ambient conditions. 123