Thermal Properties of TiO2/PbS Nanoparticle Solar Cells

Similar documents
Simulation of Type II Solar Cell by SILVACO ATLAS Software

Modified Mott-Schottky Analysis of Nanocrystal Solar Cells

Multiple Exciton Generation in Quantum Dots. James Rogers Materials 265 Professor Ram Seshadri

SUPPLEMENTARY INFORMATION

Role of Surface Chemistry on Charge Carrier Transport in Quantum Dot Solids

Colloidal quantum dot based solar cells: from materials to devices

Charge Extraction from Complex Morphologies in Bulk Heterojunctions. Michael L. Chabinyc Materials Department University of California, Santa Barbara

A Donor-Supply Electrode (DSE) for Colloidal Quantum Dot Photovoltaics

Quantum confined nanocrystals and nanostructures for high efficiency solar photoconversion Matthew C. Beard

Theoretical Study on Graphene Silicon Heterojunction Solar Cell

Nanostructured Semiconductor Crystals -- Building Blocks for Solar Cells: Shapes, Syntheses, Surface Chemistry, Quantum Confinement Effects

DEVICE CHARACTERIZATION OF (AgCu)(InGa)Se 2 SOLAR CELLS

Supplementary Figure 1. Supplementary Figure 1 Characterization of another locally gated PN junction based on boron

Solar Cell Materials and Device Characterization

Supplementary Figure S1 TEM images of a synthesis batch of PbS and Bi-doped PbS QDs (Bi/Pb=3.2%) and corresponding size distribution histograms (100

PHOTOVOLTAICS Fundamentals

All-Inorganic Colloidal Quantum Dot Photovoltaics Employing Solution-Phase Halide Passivation

UC Santa Cruz UC Santa Cruz Electronic Theses and Dissertations

Temperature Dependent Current-voltage Characteristics of P- type Crystalline Silicon Solar Cells Fabricated Using Screenprinting

Novel High-Efficiency Crystalline-Si-Based Compound. Heterojunction Solar Cells: HCT (Heterojunction with Compound. Thin-layer)

Colloidal quantum dots (CQDs) are nanoparticles, Colloidal Quantum Dot Photovoltaics: The Effect of Polydispersity

S olution processability coupled with potential for multiple exciton generation processes1 5 makes nanocrystal

Vikram Kuppa School of Energy, Environmental, Biological and Medical Engineering College of Engineering and Applied Science University of Cincinnati

Silicon quantum dot solar cell using top-down approach

The role of surface passivation for efficient and photostable PbS quantum dot solar cells

Photoelectrochemical characterization of Bi 2 S 3 thin films deposited by modified chemical bath deposition

Electroluminescence from Silicon and Germanium Nanostructures

Opto-electronic Characterization of Perovskite Thin Films & Solar Cells

High efficiency MAPbI3-xClx perovskite solar cell via interfacial passivation

Efficiency Enhancement in Polymer Solar Cell Using Solution-processed Vanadium Oxide Hole Transport Layer

Efficient Inorganic Perovskite Light-Emitting Diodes with Polyethylene Glycol Passivated Ultrathin CsPbBr 3 Films

10.6% Certified Colloidal Quantum Dot Solar Cells via Solvent-Polarity-Engineered Halide Passivation

University of Groningen. Photophysics of nanomaterials for opto-electronic applications Kahmann, Simon

Quantum Dots for Advanced Research and Devices

The broadband solar spectrum demands that solar cells be

(Co-PIs-Mark Brongersma, Yi Cui, Shanhui Fan) Stanford University. GCEP Research Symposium 2013 Stanford, CA October 9, 2013

ET3034TUx Utilization of band gap energy

Supplementary Figure S1. Verifying the CH 3 NH 3 PbI 3-x Cl x sensitized TiO 2 coating UV-vis spectrum of the solution obtained by dissolving the

As our population continues to grow, I believe that efficiently harnessing clean, abundant solar energy

Organic Electronic Devices

Numerical model of planar heterojunction organic solar cells

Graphene photodetectors with ultra-broadband and high responsivity at room temperature

Title: Ultrafast photocurrent measurement of the escape time of electrons and holes from

A quantitative model for charge carrier transport, trapping and recombination in nanocrystal-based solar cells

Mesoporous titanium dioxide electrolyte bulk heterojunction

Fundamentals of Photovoltaics: C1 Problems. R.Treharne, K. Durose, J. Major, T. Veal, V.

Colloidal PbS Quantum Dot Solar Cells with High Fill Factor

Large Storage Window in a-sinx/nc-si/a-sinx Sandwiched Structure

Thermally Stable Silver Nanowires-embedding. Metal Oxide for Schottky Junction Solar Cells

Temperature dependent behavior of lead sulfide quantum dot solar cells and films.

A Novel TiO x Protection Film for Organic Solar Cells

SUPPLEMENTARY INFORMATION

Available online at Energy Procedia 00 (2009) Energy Procedia 2 (2010) E-MRS Spring meeting 2009, Symposium B

Title of file for HTML: Supplementary Information Description: Supplementary Figures and Supplementary References

PRESENTED BY: PROF. S. Y. MENSAH F.A.A.S; F.G.A.A.S UNIVERSITY OF CAPE COAST, GHANA.

Monolayer Semiconductors

This article appeared in a journal published by Elsevier. The attached copy is furnished to the author for internal non-commercial research and

February 1, 2011 The University of Toledo, Department of Physics and Astronomy SSARE, PVIC

Solar Cells Based on. Quantum Dots: Multiple Exciton Generation and Intermediate Bands Antonio Luque, Antonio Marti, and Arthur J.

Available online at ScienceDirect. Energy Procedia 92 (2016 ) 24 28

UV Degradation and Recovery of Perovskite Solar Cells

Title: Colloidal Quantum Dots Intraband Photodetectors

Chapter 3 The InAs-Based nbn Photodetector and Dark Current

Supporting Information for Charge Generation and Energy Transfer in Hybrid Polymer/Infrared Quantum Dot Solar Cells

Novel Inorganic-Organic Perovskites for Solution Processed Photovoltaics. PIs: Mike McGehee and Hema Karunadasa

Toward a 1D Device Model Part 1: Device Fundamentals

Planar Organic Photovoltaic Device. Saiful I. Khondaker

2.626 / 2.627: Fundamentals of Photovoltaics Problem Set #3 Prof. Tonio Buonassisi

Supplementary Information

Towards a deeper understanding of polymer solar cells

Classification of Solids

Supporting Information for

Quantum Dot Technology for Low-Cost Space Power Generation for Smallsats

The Role of doping in the window layer on Performance of a InP Solar Cells USING AMPS-1D

Nanotechnology and Solar Energy. Solar Electricity Photovoltaics. Fuel from the Sun Photosynthesis Biofuels Split Water Fuel Cells

Photoconductive Atomic Force Microscopy for Understanding Nanostructures and Device Physics of Organic Solar Cells

Tianle Guo, 1 Siddharth Sampat, 1 Kehao Zhang, 2 Joshua A. Robinson, 2 Sara M. Rupich, 3 Yves J. Chabal, 3 Yuri N. Gartstein, 1 and Anton V.

Comparison of Ge, InGaAs p-n junction solar cell

Seminars in Nanosystems - I

Supporting information for: Semitransparent Polymer-Based Solar Cells with. Aluminum-Doped Zinc Oxide Electrodes

Chapter 3 Properties of Nanostructures

Optoelectronics and. Colloidal Quantum Dot. Photovoltaics. Cambridge GERASIMOS KONSTANTATOS EDWARD H. SARGENT. University of Toronto.

1. Depleted heterojunction solar cells. 2. Deposition of semiconductor layers with solution process. June 7, Yonghui Lee

Low-bandgap small molecules for near-infrared photovoltaic applications

Fundamentals of Nanoelectronics: Basic Concepts

Supporting information. Supramolecular Halogen Bond Passivation of Organometal-Halide Perovskite Solar Cells

Carrier Transport Mechanisms of a-gaas/ n-si Heterojunctions

Quantum Dot Solar Cells

Light Management across the Nano and Macro Lengthscales to Enhance Photovoltaic Device Performance

Nanocomposite photonic crystal devices

Colloidal semiconductor quantum dots (QDs) currently

Challenges in to-electric Energy Conversion: an Introduction

Light Emitting Diodes

Solar cells operation

Supporting Information Available:

EE 5611 Introduction to Microelectronic Technologies Fall Tuesday, September 23, 2014 Lecture 07

NUMERICAL SIMULATION OF THE IDEALITY FACTOR OF NON-IDEAL n-si/p-diamond HETEROJUNCTION DIODES

Organic Semiconductors for Photovoltaic Applications

Thermionic Current Modeling and Equivalent Circuit of a III-V MQW P-I-N Photovoltaic Heterostructure

Conjugated Polymers Based on Benzodithiophene for Organic Solar Cells. Wei You

Transcription:

ARTICLE Nanomaterials and Nanotechnology Thermal Properties of TiO2/PbS Nanoparticle Solar Cells Regular Paper Derek Padilla1, Guangmei Zhai2, Alison J. Breeze3, Daoli Zhang2, Glenn B. Alers1 and Sue A. Carter1,* 1 Department of Physics, University of California, Santa Cruz, California, USA 2 Department of Electronic Science and Technology, Huazhong University of Science and Technology, Wuhan, People s Republic of China 3 Solexant Corporation, San Jose, California, USA * Corresponding author E-mail: sacarter@ucsc.edu Received 13 September 2012; Accepted 10 December 2012 2012 Padilla et al.; licensee InTech. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/3.0), which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited. Abstract Photovoltaic performance is shown to depend on ligand capping on PbS nanoparticle solar cells by varying the temperature between 140K and 350K. The thermal response of open circuit voltage, short circuit current density, fill factor and shunt resistance varies between the ligands. A large increase in short circuit current density at low temperatures is observed for 1,2 ethanedithiol and 3 mercaptopropionic acid and a relatively constant short circuit current density is observed for the stiffer 1,4 benzenedithiol. Dark data provide evidence for tunnelling transport being the dominant charge conduction mechanism for all three ligand devices with recombination occurring within deep trap states. Under illumination, devices exhibit band to band recombination, indicated by an ideality factor of nearly unity. Keywords Temperature, Charge Transport, Quantum Dot, Short Circuit Current, Photovoltaic 1. Introduction Thin film quantum dot (QD) photovoltaics provide the potential to create high efficiency devices while simultaneously achieving a low manufacturing cost due to reduced material costs and less expensive processing. The observation of multiple exciton generation observed in PbS QD solutions at room temperature [1] motivates further research into such devices. Although much effort is focused on improving device performance through overall structure and QD properties, relatively few studies probe the effects of temperature or capping ligands on the photovoltaic (PV) parameters [2 5]. In solid state applications, ligands are chosen to cap QDs based on their ability to effectively transfer energy within nanostructures as well as preventing the aggregation of nanoparticles. Thiols with short chains have demonstrated both properties, making them ideal candidates for use in PV devices [6]. Previous results show that cooling the PbS QD devices through the melting point of the ligand improves performance greatly through a six fold increase in short circuit current density (Jsc) at low temperatures [7]. This was attributed to the contraction of ligands, which increased the coupling between QDs, allowing for higher current densities. In addition to the ligandsʹ thermal response within devices, the current voltage temperature parameter space reveals information regarding charge transport mechanisms. Derek Padilla, Guangmei Zhai, Alison J. Breeze, Daoli nanotechnol., Zhang, Glenn B. AlersVol. and2,sue Carter: Nanomater. 2012, Art.A.18:2012 1

Here, we present the effects of varying temperatures with three ligands used in PbS QD photovoltaic devices: 1,2 ethanedithiol (EDT), 3 mercapopropionic acid (MPA) and 1,4 benzenedithiol (BDT), by analysing current density voltage (J V) curves in dark and illuminated conditions. EDT, MPA and BDT were chosen based on their widely varying melting points (232, 291K and 370K, respectively) and their comparatively higher open circuit voltage and short circuit currents compared to longer chain ligands We show that the increase in Jsc occurs upon cooling and varies with the choice of capping ligand. In addition, MPA devices exhibit dramatically different thermal responses, as seen in the fill factor (FF) and shunt resistance (Rsh) data. came closest to matching the solar spectrum in the region of interest (1000nm to 400nm) and exhibited no changes in intensity with temperature over the temperature region used in this study. The intensity was adjusted to match the Jsc of a commercial Si device under AM1.5, 100 mw/cm2 illumination. After being fabricated in the nitrogen environment of the glove box, the devices were exposed to air for approximately five minutes while being transferred into the cryostat. Previous results show that this length of air exposure has a slightly positive effect on overall device performance [8]. While in the cryostat, nitrogen gas flowed across the devices. A RTD sensor placed next to the sample recorded the device temperature. Samples are warmed from the lowest temperature at a rate of ~15K/hr, followed by cooling at the same rate. For clarity, only cooling data is shown, with no appreciable hysteresis existing between cooling and warming data. 3. Results and Discussion The transport mechanism of the devices was determined by fitting dark J V data to the non ideal diode equation, exp, Figure 1. The exponential pre factor of each ligand device obtained through fitting Eq. (1) to dark current voltage data taken at different temperatures. Inset: Schematic of the TiO2/PbS solar cell structure. (1) where J0 is the saturation current density and A is the exponential pre factor. Both diode parameters in Eq. (1), J0 and A can be temperature dependent, with the transport mechanism being determined by the way in which they vary with temperature [9]. Figure 1 demonstrates the weak temperature dependence of A as plotted against 1/kT. This, together with the negatively sloping linear behaviour of J0 in the Arrhenius plot of Figure 2, is characteristic evidence of tunnelling transport [10]. 2. Experiments The lead sulphide (PbS) quantum dots (QDs) were provided by Solexant and the titanium oxide (TiO2) nanoparticles were purchased from Solaronix. The fabrication process follows the procedure published elsewhere [7] with the device structure shown in the inset of Figure 1. All devices use 3.9nm diameter PbS quantum dots, which correspond to an exciton peak of 1003nm and an optical energy gap of 1.24eV. Before the dipping process, the PbS QDs are submerged in a solution of chloroform and hexane with oleic acid as the capping ligand. The ligand exchange replaces oleic acid with the shorter ligands of EDT, MPA, or BDT for use in the devices (cf. Figure 2 inset). Current density voltage (J V) curves were taken from 1 to +1 V at varying temperatures in an Oxford cryostat with a Keithley 2400 source meter used a xenon light bulb next to the devices as the light source. The Xe bulb was chosen because it 2 Nanomater. nanotechnol., 2012, Vol. 2, Art. 18:2012 Figure 2. Arrhenius plot of the saturation current density for each ligand showing the negatively sloping linear behaviour. Inset: The chemical structures of the three capping ligands studied herein.

Figure 3 shows dark and light current density versus voltage (J V) data for an EDT device at a range of temperatures from 166K to 324K. The dark current density at a positive bias of 1V shows two orders of magnitude increase from low to high temperatures, indicating that thermally excited charge carriers play an important role in conduction as T increases [11]. There is no appreciable distinction between dark J V data from devices fabricated with the different ligands studied herein. difference in quasi Fermi voltages between the TiO2 and PbS layers. The Vbi for EDT, MPA and BDT devices are 0.59±0.02V, 0.56±0.03V and 0.57±0.02V respectively, indicating no significant ligand dependence of Vbi. Figure 4. Open circuit voltage dependence on temperature for three ligand devices. Extrapolating the flatter, low temperature region yields built in potential values for EDT, MPA and BDT devices (see text). The short circuit current (Jsc) plotted as a function of temperature is shown in Figure 5. As with Voc, both EDT and MPA exhibit a sharper change in Jsc than BDT. Both parametersʹ weaker dependence on temperature for BDT devices may be explained by BDTʹs more rigid structure. Comparing high temperature to low temperature Jsc values, the absolute change is similar for both EDT and MPA, however MPA shows an approximately eight fold increase while EDT is closer to two fold. BDT is relatively flat through the temperature range, though the shallow increase upon warming from below room temperature is consistent with previous results [13]. Under illumination, the ligand dependence of the devices is apparent in the PV parametersʹ temperature variations. Figure 4 shows the open circuit voltage (Voc) at a range of temperatures for devices using the three ligands. Voc is expected to vary linearly with temperature [12]. However, a clear change in slope occurs when passing through room temperature for the EDT and MPA devices. BDT devices show the most linear Voc temperature dependence. The small change that does occur is seen at higher temperatures than in the EDT and MPA devices. In FF temperature data for the three ligand devices, both EDT and BDT are relatively temperature independent (around 25%), while MPA sharply rises above room temperature from less than 20% to more than 35%. This is markedly different than the ligand dependence shown in both Jsc and Voc temperature data. The steep increase in FF occurs with a simultaneous rise in shunt resistance for MPA devices, as seen in Figure 6. Both of these together, however, do not lead to better device performance at high temperatures, mainly due to the low Jsc and Voc values. This is demonstrated in Figure 7, showing J V data for all three ligands at both 185K and 315K. The large Rsh for the MPA device at a high temperature is clearly the poorest performing J V curve. It is also seen in Figure 7 that the device made with EDT demonstrates the best EQE at low temperatures as well as a significant thermal shift compared to MPA and, especially, BDT. Extrapolation of Voc T data to 0K provides the built in potential (Vbi) of the devices [12], which indicates the The discrepancy between dark and light current densitiesʹ response to cooling, as seen in the decrease in Figure 3. Current density voltage data of an EDT device taken at 166, 190, 223, 260, 298 and 324K. (a) Dark J V measurements. (b) J V measurements under illumination. (Absolute temperatures are indicated on each J V curve.) Derek Padilla, Guangmei Zhai, Alison J. Breeze, Daoli Zhang, Glenn B. Alers and Sue A. Carter: 3

positive bias current density in Figure 3 and the increase in Jsc in Figure 5, can be reconciled by considering the weaker role of deep trap states under illumination. This can be verified by measuring an ideality factor of unity, indicating a direct, band to band recombination mechanism dominating transport, rather than recombination with charges in deep trap states. Attempting to obtain ideality factors for the devices using the dark current density from Eq. (1) yields non unity values across the temperature range, leading us to believe that deep trap states play an important role in dark charge transport, which was previously seen to inhibit charge transport in PbS devices [14]. However, when examining devices under illumination, one may look at Voc variations with light intensity, which should follow a linear trend on a semilogarithmic plot, with the slope being the ideality factor in units of thermal voltage [15]. We obtain a near unity ideality factor of 1.01 when using this technique, indicating that deep traps no longer play a significant role under illumination, which can be explained by the increased number of charge carriers filling the traps leaving mainly direct recombination processes when illuminated [15]. Figure 5. The variation of short circuit current density with temperature for the three ligand devices. Vertical dashed lines show the melting point of each ligand. Jsc dependence on temperature can partly be explained by the ligands freezing upon cooling. Previous studies have shown that ligand contraction in CdSe NP films lead to an order of magnitude decrease in transient times [16]. By varying the device ligand, we are able to verify that the freezing of the ligand does play a role in the PV properties of devices; however, not as clearly as this effect would suggest. With MPAʹs melting point being near room temperature (291K), much higher than EDT at 232K and BDTʹs being significantly higher than that at 370K, we should observe Jsc peaks at different temperatures based on which ligand is being used. This was not the case, as seen in Figure 5, although BDT does display significantly different behaviour. This difference may be attributed to the presence of a stiff benzene ring rather 4 Nanomater. nanotechnol., 2012, Vol. 2, Art. 18:2012 than the loose carbon chains found in EDT and MPA (cf. Figure 2 inset). The stability of the ring translates to much smaller changes in device performance upon cooling since the ligand is less susceptible to temperature variations. Figure 6. The shunt resistance dependence on temperature, showing MPA devicesʹ large increase above room temperature. Figure 7. J V characteristics for the photovoltaic devices fabricated with each of the three different ligands measured in the cryostat under illumination at both low temperature (185K) and high temperature (315K). 5. Conclusion In summary, we have measured the effect of capping ligands in PbS NP ultrathin film photovoltaic devices at varying temperatures through current voltage measurements. Short circuit current density and open circuit voltage show similar variations with temperature for EDT and MPA devices, both exhibiting an increase upon cooling below room temperature. Jsc and Voc for BDT devices are relatively temperature independent, which we attribute to the stiffness of the benzene ring compared to linear carbon chains. MPA devices

demonstrate a sharp increase in fill factor and shunt resistance upon warming through room temperature, however device performance is not improved due to the poor fill factor of MPA devices at warmer temperatures. The increase in dark current density upon warming is attributed to the significant role played by trap states, while light current density decreasing with higher temperatures is caused by the ligand contraction and saturation of trap states by the increased density of charge carriers when illuminated. These results indicate that even higher performance should be achievable through the careful selection of the coupling ligands. 6. Acknowledgments D. Padilla acknowledges support from a Cota Robles Fellowship and S. A. Carter acknowledges support from DOE Solar program award #1035478. 7. References [1] Ellingson R J, Beard M C, Johnson J C, Yu P, Micic O I, Nozik A J, Shabaev A, Efros A L (2005) Highly Efficient Multiple Exciton Generation in Colloidal PbSe and PbS Quantum Dots. Nano Lett. 5: 865 871. [2] Bakueva L, Musikhin S, Hines M A, Chang T W F, Tzolov M, Scholes G D, Sargent E H (2003) Size tunable infrared 1000 1600 nm electroluminescence from PbS quantum dot nanocrystals in a semiconducting polymer. Appl. Phys. Lett. 82: 2895 2897. [3] Jeong K S, Tang J, Liu H, Kim J, Schaefer A W, Kemp K, Levina L, Wang X, Hoogland S, Debnath R, Brzozowski L, Sargent E H, Asbury J B (2012) Enhanced Mobility Lifetime Products in PbS Colloidal Quantum Dot Photovoltaics. ACS Nano 6: 89 99. [4] Koleilat G I, Levina L, Shukla H, Myrskog S H, Hinds S, Pattantyus Abraham A G, Sargent E H (2008) Efficient, Stable Infrared Photovoltaics Based on Solution Cast Colloidal Quantum Dots. ACS Nano 2: 833 840. [5] Luther J M, Gao J, Lloyd M T, Semonin O E, Beard M C, Nozik A J (2010) Stability Assessment on a 3% Bilayer PbS/ZnO Quantum Dot Heterojunction Solar Cell. Adv. Mater. 22: 3704 3707. [6] Hammer N I, Emrick T, Barnes M D (2007) Quantum dots coordinated with conjugated organic ligands: new nanomaterials with novel photophysics. Nanoscale Res. Lett. 2: 282 290. [7] Ju T, Graham R L, Zhai G, Rodriguez Y W, Breeze A J, Yang L, Alers G B, Carter S A (2010) High efficiency mesoporous titanium oxide PbS quantum dot solar cells at low temperature. Appl. Phys. Lett. 97: 043106. [8] Zhai G, Bezryadina A, Breeze A J, Zhang D, Alers G B, Carter S A (2011) Air stability of TiO2/PbS colloidal nanoparticle solar cells and its impact on power efficiency. Appl. Phys. Lett. 99: 063512. [9] Marsal L F, Pallares J, Correig X, Orpella A, Bardés D, Alcubilla A (1999) Analysis of conduction mechanisms in annealed n SiC: H/p crystalline Si heterojunction diodes for different doping concentrations. J. Appl. Phys. 85: 1216 1221. [10] Park S, Cho E, Song D, Conibeer G, Green M A (2009) n Type silicon quantum dots and p type crystalline silicon heteroface solar cells. Sol. Energy Mater. Sol. Cells 93: 684 690. [11] Kundu S, Roy S K, Banerji P (2011) GaAs metal oxide semiconductor device with titanium dioxide as dielectric layer: effect of oxide thickness on the device performance. J. Phys. D: Appl. Phys. 44: 155104. [12] Hegedus S S, Shafarman W N (2004) Thin film solar cells: device measurements and analysis. Prog. Photovolt: Res. Appl. 12: 155 176. [13] Szendrei K, Speirs M, Gomulya W, Jarzab D, Manca M, Mikhnenko O V, Yarema M, Kooi B J, Heiss W, Loi M A (2012) Exploring the Origin of the Temperature Dependent Behavior of PbS Nanocrystal Thin Films and Solar Cells. Adv. Funct. Mater. 22: 1598 1605. [14] Barkhouse D A R, Pattantyus Abraham A G, Levina L, Sargent E H (2008) Thiols Passivate Recombination Centers in Colloidal Quantum Dots Leading to Enhanced Photovoltaic Device Efficiency. ACS Nano 2: 2356 2362. [15] Wetzelaer G A H, Kuik M, Lenes M, Blom P W M (2011) Determination of the trap assisted recombination strength in polymer light emitting diodes. Appl. Phys. Lett. 99: 152506. [16] Loef R, Houtepen A J, Talgorn E, Schoonman J, Goossens A (2009) Temperature Dependence of Electron Transport in CdSe Quantum Dot Films. J. Phys. Chem. C 113: 15992 15996. Derek Padilla, Guangmei Zhai, Alison J. Breeze, Daoli Zhang, Glenn B. Alers and Sue A. Carter: 5