Thioacetic-Acid Capped PbS Quantum Dot Solids Exhibiting ...

9
Thioacetic-Acid Capped PbS Quantum Dot Solids Bull. Korean Chem. Soc. 2014, Vol. 35, No. 2 457 http://dx.doi.org/10.5012/bkcs.2014.35.2.457 Thioacetic-Acid Capped PbS Quantum Dot Solids Exhibiting Thermally Activated Charge Hopping Transport Tung Duy Dao, Mahmoud Elsayed Hafez, †,‡ I. S. Beloborodov, § and Hyun-Dam Jeong †,* Department of Chemistry, Chonnam National University, Gwangju 500-757, Korea. * E-mail: [email protected] Department of Chemistry, Faculty of Science, Beni-Suef University, Beni-Suef 65211, Egypt § Department of Physics and Astronomy, California State University Northridge, Northridge, California 91330, USA Received October 4, 2013, Accepted November 13, 2013 Size-controlled lead sulfide (PbS) quantum dots were synthesized by the typical hot injection method using oleic acid (OA) as the stabilizing agent. Subsequently, the ligand exchange reaction between OA and thioacetic acid (TAA) was employed to obtain TAA-capped PbS quantum dots (PbS-TAA QDs). The condensation reaction of the TAA ligands on the surfaces of the QDs enhanced the conductivity of the PbS-TAA QDs thin films by about 2-4 orders of magnitude, as compared with that of the PbS-OA QDs thin films. The electron transport mechanism of the PbS-TAA QDs thin films was investigated by current-voltage (I-V) measurements at different temperatures in the range of 293 K-473 K. We found that the charge transport was due to sequential tunneling of charge carriers via the QDs, resulting in the thermally activated hopping process of Arrhenius behavior. Key Words : PbS quantum dot, Quantum dot solid, Thioacetic-Acid, Thin film, Charge transport Introduction Recently, among metal chalcogenide quantum dots (QDs), PbS QDs have received much attention due to their techno- logically promising optical and electronic properties. 1-3 These properties vary not only due to the quantum confinement effect, which significantly depends on the size of the PbS QDs, but also by the tunneling effect between adjacent QDs, which depends strongly on the capping ligand on the QDs. These two effects play important roles in the application of PbS QDs. 4-7 With these properties, QDs can be used in light emitting devices, such as lasers for telecommunication, 5,8,9 in modulators for extended telecommunication 10 and in infrared light emitting diodes. 11-13 In addition, due to the strong quantum size effect, PbS QDs are widely used in devices such as solar cells, 14-17 infrared detectors, 6,18 and optical switches. 19 Moreover, the multiple excitons gene- rated in PbS QDs can be detected and applied to produce highly efficient photovoltaic devices. 20,21 Quantum dot (QD) solids, which comprise an important class of artificial solids, are predicted to exhibit novel pro- perties arising from the coherent quantum mechanical inter- actions of their constituent quantum dots. 22 The characteri- stics of individual quantum dots as well as the many-body exchange interaction in quantum dot solids can be controlled to yield a new type of condensed matter. In many optoelec- tronic applications, the conductivity of the QD solids is affected by the interdistance between the QDs, that is depen- dent on the capping molecule. 23,24 Especially, PbS QDs can be prepared in many ways with different capping ligands. 25-31 Since a long and insulating ligand strongly limits the charge transport among the QDs, it was replaced with a short and more conductive one by the ligand exchange process, effec- tively enhancing electron and energy transfers and increas- ing the electronic coupling between QDs. 24 Especially, a short ligand which can undergo a condensation reaction on the surfaces of QDs can provide high electron mobility. In our previous study, under thermal annealing at 200-300 o C, the thioacetate (TAA) groups on the surfaces of cadimium sulfide (CdS) QDs condensed with the remaining L 2 Cd- (S(CO)CH 3 ) 2 (L=3,5-lutidine) molecules or with the QDs themselves to form a QD solid, which showed very high field effect mobility (48 cm 2 V 1 S 1 ) in the Al/CdS QD film/ ZrO 2 TFT structure. 32 This phenomenon was the origin of our condensable QD solids concept, which was one more clearly demonstrated in our other previous study about condensable InP QD solids. 33 Herein, the condensation pro- cess between the thioacetate groups on the surfaces of ad- jacent QDs formed an –S– linkage that provided a very dense solid. This very short interdistance of about 4.15 Å (In-S-In bond length) between the QDs by the –S– linkage is interpreted to improve greatly the electronic coupling and dipole-dipole interaction between QDs. 22,33-36 In this study, we demonstrate once again the condensable QD solids concept for the PbS QD solid, as in our previous studies about the CdS and InP QD solids. 32,33 The ligand exchange process between the long insulating oleic acid (OA) ligand and the short thioacetic acid (TAA) ligand is successfully performed by directly treating spin-coated PbS- OA thin films with TAA solutions, in order to synthesize thioacetic capped PbS QDs (PbS-TAA QDs) thin films to realize PbS QD solids. The optical and electrical properties of the condensable PbS QD solid thin films are examined to discuss the electron transport mechanism of the solids.

Transcript of Thioacetic-Acid Capped PbS Quantum Dot Solids Exhibiting ...

Page 1: Thioacetic-Acid Capped PbS Quantum Dot Solids Exhibiting ...

Thioacetic-Acid Capped PbS Quantum Dot Solids Bull. Korean Chem. Soc. 2014, Vol. 35, No. 2 457http://dx.doi.org/10.5012/bkcs.2014.35.2.457

Thioacetic-Acid Capped PbS Quantum Dot Solids Exhibiting Thermally Activated Charge Hopping Transport

Tung Duy Dao,† Mahmoud Elsayed Hafez,†,‡ I. S. Beloborodov,§ and Hyun-Dam Jeong†,*

†Department of Chemistry, Chonnam National University, Gwangju 500-757, Korea. *E-mail: [email protected]‡Department of Chemistry, Faculty of Science, Beni-Suef University, Beni-Suef 65211, Egypt

§Department of Physics and Astronomy, California State University Northridge, Northridge, California 91330, USAReceived October 4, 2013, Accepted November 13, 2013

Size-controlled lead sulfide (PbS) quantum dots were synthesized by the typical hot injection method usingoleic acid (OA) as the stabilizing agent. Subsequently, the ligand exchange reaction between OA and thioaceticacid (TAA) was employed to obtain TAA-capped PbS quantum dots (PbS-TAA QDs). The condensationreaction of the TAA ligands on the surfaces of the QDs enhanced the conductivity of the PbS-TAA QDs thinfilms by about 2-4 orders of magnitude, as compared with that of the PbS-OA QDs thin films. The electrontransport mechanism of the PbS-TAA QDs thin films was investigated by current-voltage (I-V) measurementsat different temperatures in the range of 293 K-473 K. We found that the charge transport was due to sequentialtunneling of charge carriers via the QDs, resulting in the thermally activated hopping process of Arrheniusbehavior.

Key Words : PbS quantum dot, Quantum dot solid, Thioacetic-Acid, Thin film, Charge transport

Introduction

Recently, among metal chalcogenide quantum dots (QDs),PbS QDs have received much attention due to their techno-logically promising optical and electronic properties.1-3 Theseproperties vary not only due to the quantum confinementeffect, which significantly depends on the size of the PbSQDs, but also by the tunneling effect between adjacent QDs,which depends strongly on the capping ligand on the QDs.These two effects play important roles in the application ofPbS QDs.4-7 With these properties, QDs can be used in lightemitting devices, such as lasers for telecommunication,5,8,9

in modulators for extended telecommunication10 and ininfrared light emitting diodes.11-13 In addition, due to thestrong quantum size effect, PbS QDs are widely used indevices such as solar cells,14-17 infrared detectors,6,18 andoptical switches.19 Moreover, the multiple excitons gene-rated in PbS QDs can be detected and applied to producehighly efficient photovoltaic devices.20,21

Quantum dot (QD) solids, which comprise an importantclass of artificial solids, are predicted to exhibit novel pro-perties arising from the coherent quantum mechanical inter-actions of their constituent quantum dots.22 The characteri-stics of individual quantum dots as well as the many-bodyexchange interaction in quantum dot solids can be controlledto yield a new type of condensed matter. In many optoelec-tronic applications, the conductivity of the QD solids isaffected by the interdistance between the QDs, that is depen-dent on the capping molecule.23,24 Especially, PbS QDs canbe prepared in many ways with different capping ligands.25-31

Since a long and insulating ligand strongly limits the chargetransport among the QDs, it was replaced with a short and

more conductive one by the ligand exchange process, effec-tively enhancing electron and energy transfers and increas-ing the electronic coupling between QDs.24 Especially, ashort ligand which can undergo a condensation reaction onthe surfaces of QDs can provide high electron mobility. Inour previous study, under thermal annealing at 200-300 oC,the thioacetate (TAA) groups on the surfaces of cadimiumsulfide (CdS) QDs condensed with the remaining L2Cd-(S(CO)CH3)2 (L=3,5-lutidine) molecules or with the QDsthemselves to form a QD solid, which showed very highfield effect mobility (48 cm2 V1 S1) in the Al/CdS QD film/ZrO2 TFT structure.32 This phenomenon was the origin ofour condensable QD solids concept, which was one moreclearly demonstrated in our other previous study aboutcondensable InP QD solids.33 Herein, the condensation pro-cess between the thioacetate groups on the surfaces of ad-jacent QDs formed an –S– linkage that provided a verydense solid. This very short interdistance of about 4.15 Å(In-S-In bond length) between the QDs by the –S– linkage isinterpreted to improve greatly the electronic coupling anddipole-dipole interaction between QDs.22,33-36

In this study, we demonstrate once again the condensableQD solids concept for the PbS QD solid, as in our previousstudies about the CdS and InP QD solids.32,33 The ligandexchange process between the long insulating oleic acid(OA) ligand and the short thioacetic acid (TAA) ligand issuccessfully performed by directly treating spin-coated PbS-OA thin films with TAA solutions, in order to synthesizethioacetic capped PbS QDs (PbS-TAA QDs) thin films torealize PbS QD solids. The optical and electrical propertiesof the condensable PbS QD solid thin films are examined todiscuss the electron transport mechanism of the solids.

Page 2: Thioacetic-Acid Capped PbS Quantum Dot Solids Exhibiting ...

458 Bull. Korean Chem. Soc. 2014, Vol. 35, No. 2 Tung Duy Dao et al.

Experimental

Materials. All reagents were supplied by Aldrich ChemicalCo. (St. Louis, MO, USA): lead (II) acetate trihydrate(Pb(Ac)2·3H2O), oleic acid (OA), octadecene (ODE, 90.0%), bis(trimethylsilyl)sulfide ((TMS)2S), thioacetic acid(TAA, 96%), dimethylsufoxide (DMSO), toluene (anhydrous,99.8%), and acetone.

Synthesis of PbS-OA QDs. PbS-OA QDs were synthe-sized and purified using a previously known procedure13

with a modification. First, 153 mg (0.4 mmol) lead (II) acetatetrihydrate (Pb(Ac)2·3H2O) was dissolved in a mixture of0.25 mL oleic acid (OA) and 3.75 mL octadecene (ODE,90.0%). The mixed solution was heated at 150 oC for 1 hunder vacuum condition to form the lead oleate solution,and, subsequently, the temperature was decreased to 90 oCunder argon environment. A mixture of 42 µL of bis (tri-methylsilyl) sulfide ((TMS)2S) and 2 mL of ODE wereinjected into the lead oleate solution under vigorous stirring,and the reaction was continued for nanocrystal growth for 2min. The synthesized OA-capped PbS quantum dots (PbS-OA QDs) were isolated by precipitation with acetone, thencentrifuged and redispersed in hexane. The isolation processwas repeated three times to remove the unreacted species.

The Size Control of PbS-OA QDs: The size of the PbSQDs was controlled by changing the injection temperaturefrom 90 oC to 150 oC. The diameter of the resultant PbS-OAQDs was calculated from the optical band gap value (in eV)by using the empirical equation developed by Iwan Moreelset al.37

The Fabrication of PbS-OA QD Thin Films and PbS-TAA QD Thin Films: PbS-OA QD thin films were fabri-cated by simply spin-coating the 5 wt % PbS-OA QD hexane

solution onto substrates, where the PbS-OA QDs were thesmallest ones because they had been synthesized at theinjection temperature of 90 oC. The ligand exchange processbetween the oleic acid (OA) on the QDs of the PbS-OA thinfilms with thioacetic acid (TAA) was achieved by layer - by-layer deposition consisting of the following two steps. (i)PbS-OA QD thin films were fabricated by spin - coating 25mg/mL PbS-OA QD hexane solution onto substrates. (ii)PbS-TAA QD thin films were fabricated by dipping the PbS-OA QD thin films in 0.1% TAA solution in acetonitrile (byvolume) for 60 s, followed by dipping in the acetonitrilesolution for cleaning. The layer-by-layer deposition wasrepeated several times to obtain PbS-TAA QD thin films of40-50 nm thicknesses, as depicted in Scheme 1. All the QDthin films were cured at 80 oC to remove the remainingsolvent, and then were cured at 200 oC for 1 hour under thevacuum condition of about 102 torr.

Materials Characterization. Ultraviolet-visible (UV-vis)absorption spectra were obtained using a SCINCO S-3150spectrometer. Photoluminescence spectra were obtainedusing a He-Cd (Kimmon Electric Co., IK3501R-G, Japan)excitation source at 325 nm and a photodiode array detector(IRY1024, Princeton Instrument Co., U.S.A). Fourier-trans-form infrared spectroscopy (FT-IR) was performed using aPerkinElmer (Waltham, Massachusetts, U.S.A) spectrometerwith a resolution of 8 cm1. Thermogravimetric (TG) analysiswas conducted on a METTLER TOLEDO SDTA851e underN2 flow from room temperature to 500 oC. X-ray diffraction(XRD) spectra of the PbS QDs thin films were obtainedusing an X’Pert Pro Multi Purpose X-Ray diffractometer(PANalytical, Almelo, Nerthelands) equipped with a Cu Ksource operated at 40 kV and 30 mA. The 2 angle wasscanned from 10 to 90 degrees at an increasing rate of 2o permin and step size of 0.05o. The thicknesses of PbS-OA QDand PbS-TAA QD thin films were measured by the FieldEmission Scanning Electron Microscope JSM-700F + EDS(Oxford). Metal-insulator-metal (MIM) devices were con-structed by thermal deposition of aluminum (Al) top elec-trodes on the PbS QD thin films using a shadow mask. ThePbS QDs thin films were prepared on a p-type silicon wafer,whose resistivity was less than 0.005 ·cm and thicknesswas 525 ± 25 µm. This low-resistivity silicon wafer acted asthe back electrode of the MIM devices. Current-voltage (I-V) curves were obtained in air atmosphere at differenttemperatures in the range of 293 K-473 K using a HP 4145Bsemiconductor parameter analyzer.

Results and Discussion

PbS QDs are applied in many fields mainly because oftheir tunable optical and electrical properties by variation oftheir diameter. In our study, the diameter of PbS-OA QDswas varied by changing the injection temperature. As theinjection temperature was decreased, the size of PbS quan-tum dots was decreased, as evidenced by the blue shift of theUV-vis absorption peak to a smaller wavelength (Figure1(a)). The optical band gap was calculated by Eq. (1), where

Scheme 1. Fabrication of PbS-TAA quantum dot thin film by“layer by layer” method.

Page 3: Thioacetic-Acid Capped PbS Quantum Dot Solids Exhibiting ...

Thioacetic-Acid Capped PbS Quantum Dot Solids Bull. Korean Chem. Soc. 2014, Vol. 35, No. 2 459

h is Planck’s constant, c = 3 × 108 m/s, and max is theexcitonic peak wavelength.

(1)

The diameter of the PbS-OA QDs (d in nm) was calcu-lated by using the empirical Eq. (2) developed by IwanMoreels et al.37 We found that the QD diameter was gradu-ally increased as the injection temperature was increased, asshown in Figure 1(b).

(2)

The optical properties of the smallest PbS-OA QDs, whichwas synthesized at the injection temperature of 90 oC, wereinvestigated by UV-vis absorption spectroscopy and photo-luminescence (PL) spectroscopy, as shown in Figure 2. TheUV-vis absorption and PL peaks are located at 765 and818 nm, respectively, indicating their significant blue-shiftsto the near-infrared region of 700-900 nm, as compared withthose of bulk PbS phase, whose band gap energy is 0.41 eV;max = 3020 nm. These blue shifts are surely due to thestrong quantum confinement effect.

PbS-OA QD thin films were fabricated by simply spin-

coating the 5 wt % PbS-OA QD hexane solution onto sub-strates, where the PbS-OA QDs were the smallest ones thathad been synthesized at the injection temperature of 90 oC.The ligand exchange process between the oleic acid (OA) onthe QDs in the PbS-OA thin films with thioacetic acid (TAA)was performed, as discussed in the experimental section.This layer-by-layer deposition process was repeated to givePbS-TAA QD thin films of 40-50 nm thicknesses, as de-picted in Scheme 1.

From the FT-IR spectra (Figure 3(a)) of the PbS-OA andPbS-TAA QD thin films, the intensities of the aliphatic C-Hstretching peaks around 2800-3000 cm1 for the PbS-TAAthin film were significantly decreased, as compared withthose for the PbS-OA thin film, confirming the replacementof the long chain OA ligand with the short chain TAAligand. In addition, two new peaks were observed at about980 cm1 and 624 cm1 in the FT-IR spectra of the PbS-TAAQD thin films, indicating the existence of the TAA mole-cules on the surfaces of the QDs. Based on the peak assign-ments in a previous study,38 we assert that the former 980cm1 peak is due to C-S stretching vibration in the thio-acetate resonance structure, which is 3-coordinated to thePbS QD surface, and the latter 624 cm1 peak is attributed toC-S stretching vibration in the –S-C(=O) bond, where thesulfur atom is covalently bonded to the Pb element of theQD surface. Furthermore, in the FT-IR spectrum of the TAAcapped PbS QD thin film, there are no peaks assigned tocarbonyl group in the range of 1700 cm1-1750 cm1. Instead,the peaks around 1400 cm1-1500 cm1 are attributed to thethioacetate (-SC(O)CH3) resonance structure which is 3-coordinated to the PbS QD surface. This kind of decrease invibrational wavenumber of the carbonyl group is similarlyobserved even in the FT-IR results of the OA-capped PbSQD thin film where the peaks of carbonyl group in the rangeof 1700 cm1-1750 cm1 are not observed and, instead of it,those of carboxylate resonance structure 3-coordinated tothe PbS QD surface in the range of 1400 cm1-1500 cm1 arefound. Moreover, the disappearance of S-H stretching at

Eg = hcmax

---------

Eg = 0.41 + 1

0.0252d2 0.283d+------------------------------------------

Figure 1. (a) UV-vis spectra of the PbS-OA QDs (in hexanesolution) for different injection temperatures and (b) Variations ofthe QDs diameter according to injection temperature.

Figure 2. UV-vis absorption and PL spectra of the PbS-OA QDs inhexane solution, which were used in ligand exchange.

Page 4: Thioacetic-Acid Capped PbS Quantum Dot Solids Exhibiting ...

460 Bull. Korean Chem. Soc. 2014, Vol. 35, No. 2 Tung Duy Dao et al.

2600-2550 cm1 was due to the bonding of the sulfur atomwith the lead atom on the QD surfaces,39 implying a reactionbetween TAA molecules and PbS QDs. Figure 3(b) shows apeak at 3005 cm1 corresponding to Csp2-H stretching vib-ration on the OA ligand. After the ligand exchange process,the peak intensity was decreased. The Csp2-H infrared ab-sorption cross section is assumed to be independent ofsurface coverage as confirmed by similar surfaces.40 There-fore, the surface coverage of the TAA ligands (or ligandexchange percentage) was determined from the relativedecrease in the integrated IR absorbance of the Csp2-H peakgiven by Eq. (3):

Ligand exchange % = (3)

where and are the integrated absorbancesdue to the Csp2-H peak at 3005 cm1 before and after theligand exchange, respectively. Finally, the surface coverageof TAA ligands (or the ligand exchange percentage) isdetermined to be 64%, implying that the surface coverage ofOA ligands continues to be 36%.

As shown in Figure 4, the XRD patterns indicate the highcrystallinity of the PbS-TAA QD thin films,showing variousdiffraction peaks at 2 values of 26.02o, 30.09o, 43.07o,50.98o, 53.36o, 62.88o, 69.02o, 70.88o and 79.80o. All thesediffraction peaks can be assigned to the diffractions from the(111), (200), (220), (311), (222), (400), (331), (420), and(422) planes, respectively, of the face-center-cubic rock saltstructured PbS.2,31,36 The average diameter of the QDs in thethin films (D) was calculated from the Debye-Scherrerformula,41

(4)

where k is the shape factor of about 0.9, is the x-ray wave-length (0.15405 nm), is the full width at half maximum(FWHM) of the main diffraction peak from the (111) planeand is the diffraction angle. The average diameters of PbS-OA QD and PbS-TAA QDs in the thin films were 3.03 nmand 3.74 nm, respectively, indicating that a surface reactionmust have taken place during the dipping step of the PbS-OAA QD thin films in the TAA solution in the ligandexchange process. The increase in the diameter of the QDscan be explained by the Osttwald ripening process: a largeparticle grows at the expense of smaller ones. This phen-omenon was previously mentioned by Kumacheva’s groupfor thiolate capped-PbS QDs.42 The TAA ligand, having ahigher electron donating ability than the OA ligand, isthought to etch away the Pb element in the PbS QD surfacelike the OA ligand,43 as shown in the upper part of Scheme2, inducing the breakage of the Pb-S bond on the QDssurfaces producing smaller quantum dots. Subsequently, thesmall QDs in the Ostwald ripening process are exhausted toproduce larger QDs, as shown in the lower part of Scheme 2.

In order to validate the etching mechanism of the surfacesof the PbS QDs by the TAA molecules, the following testligand exchange reaction in PbS-OA QDs solution wasperformed. After the TAA molecules were added to the PbS-OA QDs in hexane, a black precipitate of PbS-TAA QDs

IAC

sp2before

IAC

sp2after

IAC

sp2before

-------------------------------------------------

IACsp2

before IACsp2

after

D = k

cos--------------

Figure 3. (a) FT-IR spectra of PbS-OA and PbS-TAA QDs thinfilms, (b) FT-IR spectra of C=C bond peak of OA in PbS-OA andPbS-TAA QDs thin films.

Figure 4. X-ray diffraction (XRD) data of PbS-TAA QDs (redline) and PbS-OA QDs (blue line).

Page 5: Thioacetic-Acid Capped PbS Quantum Dot Solids Exhibiting ...

Thioacetic-Acid Capped PbS Quantum Dot Solids Bull. Korean Chem. Soc. 2014, Vol. 35, No. 2 461

was abruptly formed. The black QDs precipitate was puri-fied by centrifugation. Then, we monitored the change of theblack QDs precipitate in the DMSO solvent to confirm theetching mechanism. The precipitate was not soluble at allinitially, but after three days, the solution became darkorange, with the precipitate still existing in the solution,implying that the precipitate was partially soluble in theDMSO solvent. This result suggests that some of the QDs inthe DMSO solvent became smaller due to the etchingreaction, which generated the lead thioacetate complexPb(S(CO)CH3)x that is dissolved into the DMSO solvent.Pb(S(CO)CH3)x can dissolve in DMSO and turn the solu-tion dark orange. To confirm the color of the lead thio-acetate complex Pb(S(CO)CH3)x, we mixed lead acetatePb(OCOCH3)2 with thioacetic acid in the DMSO solvent,stirred the solution for 30 min, and then, filtered it toobtain a dark - orange solution with low concentration ofPb(S(CO)CH3)x. The UV spectrum of the PbS-TAA QDssolution, which was stirred in the DMSO solvent for 3 days,was similar to that of the lead thioacetate complexPb(S(CO)CH3)x (Supporting information Fig. S1).

The effects of the capping ligand and annealing temper-ature on the UV-vis absorption spectra of the PbS QD thinfilms are summarized in Figure 5(a). The excitonic peaks ofthe PbS-TAA QDs thin films were more red-shifted com-pared with those of the PbS-OA QDs thin films, wellconsistent with the size increase of PbS-OA QDs (3.03 nm)to the size of PbS-TAA QDs (3.74 nm) in the XRD result.The excitonic peaks of the PbS-OA QDs thin films weremaintained with increasing annealing temperature, whereasthose of the PbS-TAA QDs thin films were red-shifted from886 nm to 906 nm with increasing curing temperature from80 oC to 200 oC. We suggest that this red shift was originatedfrom the condensation reaction between the thioacetategroups on the adjacent QDs at the high annealing temper-atures (reaction R1). This condensation reaction decreasesthe distance between adjacent QDs, which was the case withcondensable InP QDs.33 We can find an another researchresult to ensure the possibility of the condensation reactionof two thioacetate groups coordinated to Pb metal ions in aprevious study,44 where Pb(SCOCH3)2·18-crown-6 undergoesthermal decomposition to give rise to crystalline PbS phase.The formation of PbS phase cannot explained chemicallywithout assuming the condensation reaction of two thio-acetate groups coordinated to the Pb metal ions.

(R1)

Nozik’s group reported a red shift for InP QDs upon UVabsorption due to the decrease in the interdistance betweenthe QDs on the thin film. The InP QDs films with the inter-distances of 0.9 nm and 1.8 nm showed red shifts of 140meV and 18 meV, respectively.23 In our previous study ofInP-TAA QDs thin films, we also observed a red-shift from550 nm to 570 nm as the curing temperature is increased.33

This shift was due to the condensation reaction that decreasesthe interdistance between the adjacent QDs. Thus, a changeon the interdistance between the QDs is the key to the redshift upon UV absorption. Two physical phenomena owingto the decrease in the interdistance can explain the red shift:(i) electronic coupling and (ii) dipole-dipole interaction.Electronic coupling is generated when the interdistancebetween the QDs is decreased to an extent that the electronicwave functions of individual QDs in close proximity canhybridize together to form delocalized states or mini-bands.23,33,36 This electronic coupling is dependent on thesize of QDs23,45-47 and is increased exponentially with thedecrease in the interdistance between the QDs. The increase

Scheme 2. Expected scheme for increasing the size of QDs duringthe ligand exchange process.

Figure 5. (a) UV-vis absorption spectra of PbS-OA and PbS-TAAQDs thin films cured at 80 oC and 200 oC, respectively, and (b) PLspectra of PbS-OA and PbS-TAA QDs thin films cured at 200 oC.

Page 6: Thioacetic-Acid Capped PbS Quantum Dot Solids Exhibiting ...

462 Bull. Korean Chem. Soc. 2014, Vol. 35, No. 2 Tung Duy Dao et al.

in the electronic coupling reduces the energy of excitonictransition, resulting in the red shift on the UV absorptionspectroscopy. In addition, dipole-dipole interaction betweenthe QDs due to the decrease in the interdistance between theQDs is also related to the red-shifts mentioned above.35 Inthis model, the transition dipole moment of light absorbingQDs may induce dipole moments in the neighboring QDs intheir ground state. The dipole-dipole interaction adds anadditional Coulomb term to the dipole transition, thus reduc-ing the energy of excitonic transition.

At the same measurement conditions, the PbS-OA QDsthin films showed high photoluminescence (PL) efficiencies,while the PbS-TAA QDs films showed complete PL quen-ching (Figure 5(b)), which can be explained by the increasednumber of surface dangling bonds after the ligand exchangeprocess. Additionally, a strong electronic coupling betweenthe neighboring quantum dots can cause hybridization ofband edge orbitals and the transfer of exciton energy from alarge bandgap quantum dot to small bandgap quantum dot,which can decrease in the PL intensity; therefore, a strongelectronic coupling is responsible for the PL quenchingphenomenon.48 In addition, the strong dipole-dipole inter-action in a PbS-TAA QDs film enhances the FRET (Försterresonance energy transfer) process inside the PbS-TAA QDs,which leads to PL quenching.33

A metal-insulator-metal (MIM) structure was fabricated toinvestigate the electrical properties of the PbS-OA QDs andthe PbS-TAA QDs thin films, whose thicknesses were thesame in the SEM images (Figure 6). After the ligand ex-change from the OA (C17H33COOH) to TAA (CH3COSH)on the surface of a PbS quantum dot, the number of carbonatoms on the alkyl chain was decreased by about 17 times,according to the equation G-G0exp(-n), where G0 is thepre-factor, is the decay constant (~1.2) and n is thenumber of carbon atoms on the alkyl chain.49 At the initialstate of baking at 80 oC, the conductivity of the PbS QDsthin films is expected to increase by about 8.3 orders ofmagnitude after the ligand exchange process:

(5)

where G1 and G2 are the conductances of the PbS-OA QDsand PbS-TAA QDs thin films.

In addition, after the ligand exchange process and afterforming the completely condensed state of curing at 200 oC,we can expect that the conductivity of the PbS-TAA thinfilms will be significantly enhanced by at least 8.3 orders of

magnitude. In reality, at 200 oC curing, the current densitiesacross the PbS-TAA QDs thin films were higher than thoseof the PbS-OA QDs thin films by just about 2-4 orders ofmagnitude, as shown in Figure 7. These differences in thecurrent density are much smaller than what was expectedfrom the above theoretical estimation. This is may be due tothe low ligand exchange percentage and the increase in theQD diameter, which were mentioned in the above resultsand discussions, and due to the increase in defect concent-ration.

For a more in-depth understanding of the charge transportmechanism in the PbS-TAA QDs thin films, as shown inFigure 8(a), we measured the I-V curves (current density-electric field curves) at different temperatures 293 K, 323 K,373 K, 423 K, and 473 K for the PbS-TAA QDs film curedat 200 oC. All of curves were symmetric, sigmoid-shaped.

We mention that the effect of the irregularities in the QDspositions and in the strengths of the tunneling coupling onthe physical properties of QD arrays is different betweenmetallic and insulating samples. If the coupling between thedots is sufficiently strong and the system is well conducting,the irregularities are not very important.50 In contrast, irregu-larities become crucial in the limit of low coupling, wherethe system is an insulator.51 The key parameter that deter-mines most of the physical properties of a QD array is thedimensional tunneling conductance g between neighboringdots.52 Samples with g > 1 exhibit metallic transport properties,while those with g < 1 show insulating behavior. All oursamples belong to the insulating side of the metal-insulatortransition. Therefore, the charge on each dot becomes quan-tized as in the standard Coulomb blockade behavior. In thiscase, an electron has to overcome an electrostatic barrier ofthe order EC in order to hop onto a neighboring dot, with EC

being the Coulomb energy of a single dot. The problem ofelectron transport in QD arrays is twofold: (i) to understandthe behavior of the density of states near the Fermi level andthe role of the Coulomb correlations in forming this densityof states, and (ii) to study the mechanism of electron tunnel-ing through a dense array of QDs.

G2

G1

------ = G0exp 1.2– 1

G0exp 1.2– 17 ----------------------------------------- = exp 19.2 = 108.3

Figure 6. FE-SEM of PbS-OA and PbS-TAA films cured at 200 oC.

Figure 7. Current density-electric field curves of PbS-OA (blueline) and PbS-TAA (red line) films cured at 200 oC. Inset pictureshows the MIM structure.

Page 7: Thioacetic-Acid Capped PbS Quantum Dot Solids Exhibiting ...

Thioacetic-Acid Capped PbS Quantum Dot Solids Bull. Korean Chem. Soc. 2014, Vol. 35, No. 2 463

Charge disorder arises from variations in the local chemicalpotentials due to polarization by trapped parasitic charges inthe substrate or in the ligand shells surrounding the dots. Inpractice, charge disorder is unavoidable for QD arrays; as aresult, tunneling occurs along paths that optimize the overallenergy cost. This results in two distinct regimes.53 At appliedbias voltages and temperatures large enough to overcomelocal Coulomb energy costs, transport occurs by sequentialtunneling between neighboring dots. At small bias and lowtemperatures, sequential tunneling is suppressed by the Cou-lomb blockade. In this regime, conduction involves higherorder tunneling processes, the so-called cotunneling events,which can transport a charge over distances of several dotswithout incurring the full Coulomb energy costs. The cross-over temperature between sequential tunneling and electroncotunneling is defined as Tcross = Ec

2/T0, where T0 = e2/ isthe characteristic temperature with and being the effec-tive dielectric constant and localization length, respectively.In our experiment, T > Tcross and therefore, electron transportwas due to sequential tunneling, which led to the Arrheniusbehavior for zero bias conductance.

Important information on charge transport mechanism canbe deduced from the relationship between conductance (G)

and temperature, where the G is simply calculated from thecurrent density. The dependence of ln(G) on 1/T for the PbS-TAA QDs thin film annealed at 200 oC is shown in Figure8(b). The values of lnG(T) show a linear relation from 323 Kto 473 K, otherwise known as an Arrhenius-like T depen-dence.54 This relation confirms the sequential tunnelingmechanism mentioned in the above discussion. The trans-port in this regime T > 323 K was a thermally activatedtransport of sequential tunneling shown in Figure 9. There-fore, Eq. (6) is as follows: 55

G(, T) = G0exp[d]exp[Ea/kBT] (6)

where Ea is the activation energy for charge transport, kB isthe Boltzmann constant, d is the electron tunneling coeffi-cient and is the QD edge-to-edge distance, in other words,interdistane. From the slope of the Arrhenius plot, weobtained the activation energy (Ea = 170 meV). In a previousreport on gold nanoparticles connected by oligothiophenemolecules,55 the activation energy was about 21-45 meV,which is smaller than our result (170 meV). Another, ourresult is similar with the previous study by Sargent’s group,who reported an activation energy about 160 meV for n-butylamine capped PbS-QDs.56 This similarity in theactivation energy is very unusual if one considers that thecondensation reaction of TAA ligands between the adjacentQDs gives Pb–S–Pb linkages of very short interdistance of4.34 Å while the n-butylamines ligands of chain length of5.12 Å provide an interdistance of approximately 10.24 Å.57

This is because a shorter interdistance give rise to a loweractivation energy. The activation energy of 170 meV for ourTAA-capped PbS QD solids, which is higher than expected,is attributed to the low ligand exchange percentage of 64%,which was previously mentioned in the FT-IR results. Theremaining OA ligands are not likely to allow the physicalsituation in which the QDs interdistance is decreased to theextent of Pb-S-Pb linkage length between the QDs.

One assumes that, as for the interdistance of the PbS-TAAQDs, 64% of the sphere surface of a QD is separated by Pb-S-Pb linker length of 4.34 Å while the other 36% spheresurface of the QD is separated by 2 times the OA chainlength of 39.2 Å as shown in Scheme 3.57 Herein, we can

Figure 8. (a) I-V curves of PbS-TAA thin films cured at 200 oC fordifferent temperatures. Inset picture shows the linear fit that allowsthe determination of the differential conductance G = dI/dV at thezero bias limit. (b) Inverse relationship between lnG and absolute T.

Figure 9. Activated hopping transport of sequential tunneling ofPbS-TAA QDs thin films.

Page 8: Thioacetic-Acid Capped PbS Quantum Dot Solids Exhibiting ...

464 Bull. Korean Chem. Soc. 2014, Vol. 35, No. 2 Tung Duy Dao et al.

assert that first-order electron hopping rate constant, kET

(s1), can be calculated from nanoparticle film conductivitiesby assuming a cubic-lattice model, which was previouslyused to examine the electron transport for redox polymersarenethiolate-protected nanoparticles,58

(7)

where, R is the gas constant; F the Faraday constant; theconductivity per unit area ( = G·thickness/area; 1cm1) attemperature T (K); the averaged core edge-to-edge distance(cm), as expressed by:

(8)

where = 0.64 (ligand exchange ratio), lPb-S and lOA are thelengths of the Pb-S bond and the oleic acid chain; and Cexpt

the solid state concentration (mol/cm3) of a PbS quantumdot, as expressed by:

(9)

where NA is Avogadro’s number and 0.52 is the filling factorfor the PbS cubically close packed film, and the 103 factorconverts the dimensions to molar concentrations, and r core isthe radius of the PbS-TAA quantum dot. The results of thefirst-order electron hopping rate constant, kET (s1), aresummarized in Table 1. The kET = 3.41 × 106 s1 for the PbS-TAA QD thin film cured at 200 ×C is estimated to be muchsmaller than that (2.6 × 108 - 1.1 × 1011 s1) of the arenethi-olate capped-Au nanoparticle thin films.58 This result sup-ports the idea that the remaining long insulating OA, evenafter the ligand exchange process, limits the electron transferin PbS-TAA QD solids.

In our PbS-TAA QDs thin film, its electrical propertiesand activation energy in the activated transport mechanismof sequential tunneling are expected to be significantlyaffected by the ligand exchange percentage and the relevant

interfacial chemical state between PbS-TAA QDs which aresurely to be tuned by changing ligand exchange reactiontime in the layer-by-layer film formation. This issue will befurther investigated with more controlled experiments in aseparate study.

Conclusion

Thioacetic acid capped-lead sulfide quantum dot (PbS-TAA QD) thin films were synthesized by the ligand ex-change process of oleic acid capped- lead sulfide quantumdot (PbS-OA QD) thin films with thioacetic acid (TAA)using the layer–by–layer method, where ligand exchangepercentage was 64%. By applying the ligand exchange pro-cess, the size of the quantum dots in the thin films wasincreased from 3.03 nm to 3.74 nm due to the Ostwald ripen-ing process. The electrical conductivity of the PbS-TAA QDthin films was enhanced by about 2-4 orders of magnitude,as compared with that of the PbS-OA QD thin films. Therm-ally activated hopping transport of sequential tunneling wasdetermined to be the dominant carrier transport mechanismin the PbS-TAA QD thin films. The activation energy ofPbS-TAA QD thin films was estimated to be about 170 meVfrom the temperature-dependent electrical conductivity,which was not much higher than the activation energy of then-butylamine capped PbS QDs thin film, because of the lowligand exchange percentage of 64% of the PbS-TAA QDthin films.

Acknowledgments. This work was also supported by theMaterials Original Technology Program (10041222) fundedby the Ministry of Knowledge Economy (MKE, Korea). I.B. was supported by NSF Grant DMR 1158666.

References

1. Justo, Y.; Goris, B.; Kamal, J. S.; Geiregat, P.; Bals, S.; Hens, Z. J.Am. Chem. Soc. 2012, 134, 5484.

2. He, X.; Demchenko, I. N.; Stolte, W. C.; Buuren, A. V.; Liang, H.J. Phys. Chem. C 2012, 116, 22001.

3. Uhuegbu; Chidi, C. Can. J. Scient. Ind. Res. 2011, 2, 230. 4. Hyun, B. R.; Zhong, Y. W.; Bartnik, A. C.; Sun, L.; Abruna, H. D.;

Wise, F. W.; Goodreau, J. D.; Matthews, J. R.; Leslie, T. M.;Borrelli, N. F. ACS NANO 2008, 2, 2206.

5. Wise, F. W. Acc. Chem. Res. 2000, 33, 773. 6. Mcdonald, S. A.; Konstantatos, G.; Zhang S.; Cyr, P. W.; Klem, E.

J. D.; Levina, L.; Sargent, E. H. Nat. Mater. 2005, 4, 138. 7. Zhang, S.; Cyr, P. W.; Mcdonald, S. A.; Konstantatos, G.; Sargent,

E. H. Appl. Phys. Lett. 2005, 87, 233101. 8. Hens, Z.; Vanmaekelbergh, D.; Stoffels, E. J. A. J.; Kempen, H. V.

Phys. Rev. Lett. 2002, 88, 236803. 9. Dantas, N. O.; Qu, F.; Silva, R. S. J. Phys. Chem. B 2002, 106,

7453.10. Klem, E. J. D.; Levina, L.; Sarget, E. H. Appl. Phys. Lett. 2005,

87, 053101.11. Bourdakos, K. N.; Dissanayake, D. M. N. M.; Lutz, T.; Silva, S.

R. P.; Curry, R. J. Appl. Phys. Lett. 2008, 92, 153311.12. Kanstantatos, G.; Huang, C.; Levina, L.; Lu, Z.; Sargent, E. H.

Adv. Funct. Mater. 2005, 15, 1865.13. Hines, M. A.; Scholes, G. D. Adv. Mater. 2003, 15, 1844.14. Gunes, S.; Fritz, K. P.; Neugebauer, H.; Sariciftci, N. S.; Kumar,

kET s 1– = 6RT

10 3– F22Cexpt

-------------------------------

= 2 lPb-S 1 – lOA+

Cexpt = 0.52 103

43--- rcore lPb-S+ 3 + 1 – rcore lOA+ 3 NA

-----------------------------------------------------------------------------------------------------

Scheme 3. One assumes that, for the interdistance of the PbS-TAAQDs, 64% of the surface of a QD is separated by the Pb-S-Pblinker bond length of 4.34 Å while the other 36% of the surface isseparated by 2 times the OA length of 39.2 Å.

Table 1. Activation energy, conductivity data and self-exchangerate constants for PbS-TAA quantum dot thin films

Ea (meV) (1cm1) at 323 K Cexpt (M) kET(s1) at 323 K

170 4.4 × 107 7.86 × 103 3.41 × 106

Page 9: Thioacetic-Acid Capped PbS Quantum Dot Solids Exhibiting ...

Thioacetic-Acid Capped PbS Quantum Dot Solids Bull. Korean Chem. Soc. 2014, Vol. 35, No. 2 465

S.; Scholes, G. D. Sol. Energy Mater. Sol. Cells 2007, 91, 420.15. Wang, P.; Wang, L.; Ma, B.; Li, B.; Qiu. Y. J. Phys. Chem. B 2006,

110, 14406.16. Bayon, R.; Musembi, R.; Belaidi, A.; Bar, M.; Guminskaya, T.;

Lux-Steiner, M. Ch.; Dittrich, Th. Sol. Energy Mater. Sol. Cells2005, 89, 13.

17. Klem, E. J. D.; MacNeil, D. D.; Cyr, P. W.; Levina, L.; Sargent, E.H. Appl. Phys. Lett. 2007, 90, 183113.

18. Gadenne, P.; Yagil, Y.; Deutscher, G. J. Appl. Phys. 1989, 66,3019.

19. Kane, R. S.; Cohen, R. E.; Silbey, R. J. Phys. Chem. 1996, 100,7928.

20. Schaller, R. D.; Sykora, M.; Pietryga, J. M.; Klimov, V. I. NanoLett. 2006, 6, 424.

21. Ellingson, R. J.; Beard, M. C.; Johnson, J. C.; Yu, P.; Micic, O. I.;Nozik, A. J.; Shabaev, A.; Efros, A. L. Nano Lett. 2005, 5, 865.

22. Hanrath, T. J. Vac. Sci. Technol. A 2012, 30, 03082. 23. Nozik, A. J.; Beard, M. C.; Luther, J. M.; Law, M.; Ellingson, R.

J.; John, J. C. Chem. Rev. 2010, 110, 6873. 24. Talapin, D. V.; Lee, J.-S.; Kovalenko, M. V.; Shevchenko, E. V.

Chem. Rev. 2010, 110, 389. 25. Qu, F.; Silva, R. S.; Dantas, N. O. Phys. Stat. Sol. B 2002, 232, 95.26. Lu, X.; Zhao, Y.; Wang, C. Adv. Mater. 2005, 17, 2485.27. Choudhury, K. R.; Sahoo, Y.; Jang, S.; Prasad, P. N. Adv. Funct.

Mater. 2005, 15, 751.28. Watt, A.; Thomsen, E.; Meredith, P.; Dunlop, H. R. Chem. Commun.

2004, 20, 2334.29. Zhou, Y.; Itoh, H.; Uemura, T.; Naka, K.; Chujo, Y. Langmuir

2002, 18, 5287.30. Wu, S.; Zeng, H.; Schelly, Z. A. Langmuir 2005, 21, 686.31. Zhang, Z.; Lee, S. H.; Vittal, J. J.; Chin, W. S. J. Phys. Chem. B

2006, 110, 6649. 32. Seon, J. B.; Lee, S.; Kim, J. M.; Jeong, H. D. Chem. Mater. 2009,

21, 604.33. Dung, M. X.; Tung, D. D.; Jeong, H. D. Curr Appl Phys. 2013, 13,

1075.34. Baik, S. J.; Kim, K.; Lim, K. S.; Jung, S.; Park, Y. C.; Han, D. G.;

Lim, S.; Yoo, S.; Jeong, S. J. Phys. Chem. C 2011, 115, 607.35. Döllefeld, H.; Weller, H.; Eychmüller, A. J. Phys. Chem. B 2002,

106, 5604.36. Miæiæ, O. I.; Ahrenkiel, S. P.; Nozik, A. J. Appl. Phys. Lett. 2001,

78, 4022.37. Moreels, I.; Lambert, K.; Smeets, D.; DeMuynk, D.; Nollet, T.;

Martins, J.; Vanhaecke, F.; Vantomme, A.; Delerue, C.; Allan, G.;Hens, Z. ACS NANO 2009, 3, 3023.

38. Deivaraj, T. C.; Lin, M.; Loh, K. P.; Yeadon, M.; Vittal, J. J. J.Mater. Chem. 2003, 13, 1149.

39. Konstantatos, G.; Howard, I.; Fischer, A.; Hoogland, S.; Clifford,J.; Klem, E.; Levina, L.; Sargent, E. H. Nature 2006, 422, 180.

40. Weeks, S. L; Macco, B.; van de Sanden, M. C. M.; Agarwal, S.Langmuir 2012, 28, 17295.

41. Zhang, D.; Song, J.; Zhang, J.; Wang, Y.; Zhanga, S.; Miao, X.CrystEngComm. 2013, 15, 2532.

42. Zhao, X.; Gerelicov, I.; Musikhin, S.; Cauchi, S.; Sukhovatkin, V.;Sargent, S. H.; Kumacheva, E. Langmuir 2005, 21, 1086.

43. Zhang, H.; Hu, B.; Sun, L.; Huvden, R.; Wise, F. W.; Muller, D.A.; Robinson, R. D. Nano Lett. 2011, 11, 5356.

44. US Patent 5837320, 1998. 45. Lee, J.; Choi, O.; Sim, E. J. Phys. Chem. Lett. 2012, 3, 714.46. Koole, R.; Liljeroth, P.; Donegá, C. M.; Vanmaekelbergh, D.;

Meijerink, A. J. Am. Chem. Soc. 2006, 128, 32.47. Leatherdale, C. A.; Kagan, C. R.; Morgan, N. Y.; Empedocles, S.

A.; Kastner, M. A.; Bawendi, M. G. Phys. Rev. B 2000, 62, 2669. 48. Noh, M.; Kim, T.; Lee, H.; Kim, C. K.; Joo, S. W.; Lee, K. Colloids

and Surface A: Physicochem. Eng. Aspects. 2010, 359, 39.49. Zabet-Khosousi, A.; Dhirani, A. D. Chem. Rev. 2008, 108, 4072.50. Beloborodov, I. S.; Efetov, K. B.; Lopatin, A. V.; Vinokur, V. M.

Phys. Rev. Lett. 2003, 91, 246801.51. Beloborodov, I. S.; Lopatin, A. V.; Vinokur, V. M. Phys. Rev. B

2005, 72, 125121.52. Beloborodov, I. S.; Efetov, K. B.; Lopatin, A. V.; Vinokur, V. M.

Reviews of Modern Physics 2007, 79, 469. 53. Beloborodov, I. S.; Glatz, A.; Vinokur, V. M. Phys. Rev. B 2007,

75, 052302.54. Rmero, H. E.; Drndic, M. Phys. Rev. Lett. 2005, 95, 156801.55. Taniguchi, S.; Minamoto, M.; Matsushita, M. M.; Sugawara, T.;

Kawada, Y.; Betthell, D. J. Mater. Chem. 2006, 16, 3459.56. Konstantatos, G.; Sargent, E. H. Appl. Phys. Lett. 2007, 91, 173505.57. The bond lengths and molecular fragment lengths were determined

from optimized structures obtained by the calculation usingGaussian 03 program.

58. Wuelfing, W. P.; Murray, R. W. J. Phys. Chem. B 2002, 106, 3139.