Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Article

pubs.acs.org/JPCC

Magnitude of the Förster Radius in Colloidal Quantum Dot Solids


A. Jolene Mork,† Mark C. Weidman,‡ Ferry Prins,‡ and William A. Tisdale*,‡

Department of Chemistry and ‡Department of Chemical Engineering, Massachusetts Institute of Technology, Cambridge,
Massachusetts 02139, United States

ABSTRACT: Excitonic energy transfer among colloidal nanocrystal quantum dots (QDs) is
responsible for exciton transport in many QD optoelectronic devices. While Förster theory has
successfully accounted for the distance scaling of energy transfer in many QD systems, the
overall magnitude of the Förster radius in close-packed QD solids remains an open question.
Here, we use spectrally resolved transient photoluminescence quenching to measure the
magnitude of the Förster radius in blended donor−acceptor QD assemblies. For blends of
CdSe/CdZnS core/shell QDs consisting of 4.0 nm diameter donors (λmax,em ≈ 550 nm) and
5.5 nm acceptors (λmax,abs ≈ 590 nm), we measure energy transfer rates per donor−acceptor
pair that are 10−100 times faster than the predictions of Förster theory. These rates
correspond to an effective Förster radius of 8−9 nm, compared to a theoretical Förster radius
of 5−6 nm. We discuss possible sources for the discrepancy between theory and experiment
including the magnitude of the absorption cross section, dipole orientation, and dipole−multipole couplingand suggest that
several common assumptions should be considered carefully before applying Förster theory to solid QD films.

■ INTRODUCTION
The unique size-dependent optical characteristics of semi-
comparatively large size of QDs has led previous authors to
question whether QDs behave as point dipoles separated by a
conductor nanocrystal quantum dots (QDs) have motivated distance equal to the center-to-center spacing, as the model
research in next-generation technologies ranging from solid- suggests.19−21,27 Prior experiments in solid films have measured
state optoelectronic devices, such as LEDs,1−3 solar cells,4,5 the energy transfer rate, kET, between donor and acceptor QDs
lasers,6,7 photodetectors,8 and luminescent solar concentrators,9 through monitoring the donor fluorescence lifetime in either
to biomedical technologies utilizing imaging, sensing, and donor−acceptor blended films25,26,28,29 or bilayer struc-
labeling with QDs.10,11 The physical properties of colloidal tures,15,26,30,31 and have calculated values for kET in both
QDs, such as broadband absorption, narrow emission line geometries on the order of (1−10) × 108 s−1. Changes in
width, band gap tunability, and high quantum yield, have deposition technique (drop-casting vs Langmuir−Blodgett)15
allowed commercialization of QD technologies;3,12 for many of as well as changes in the spacing between donor and acceptor
these technologies, energy transfer between QDs must be layers in bilayer structures30 or with a ligand series32 have been
controlled to achieve optimum performance. For example, shown to affect kET by altering the proximity of nearest acceptor
white-light-emitting QD-LEDs require an excess of blue neighbors. Similarly, for energy transfer within an inhomoge-
emitters relative to green and red in order to compensate for neous distribution of QDs, the line width and Stokes shift of
energy transfer to, and emission from, smaller band gap QDs affect the energy transfer rate from large band gap to small
(redder) dots.13 Similarly, control of exciton diffusion14 via band gap QDs, consistent with the expected dependence of
cascaded energy transfer structures allows for efficient exciton spectral overlap on the energy transfer rate predicted by Förster
funneling to acceptor sites and dramatically increases the theory.33,34 These experiments have revealed that Förster
fluorescence intensity of the acceptor.15,16 Several biological theory correctly predicts some scaling parameters that affect the
detection assays rely on energy transfer from QDs to organic energy transfer rate in QD solids, such as overlap integral and
chromophores, where the proximity of a donor and acceptor interparticle separation. However, few have made quantitatively
can be inferred from the energy transfer efficiency.17,18 Förster rigorous comparison to numerical predictions based on spectral
resonance energy transfer (FRET) theory has been well properties in order to test the limits of Förster theory as it
established as a theoretical and practical framework for applies to solid QD films.
understanding energy transfer in QD systems, in particular In this work, we measure energy transfer rates from 4.0 nm
between QDs and organic chromophores in solution.19−22 diameter donor QDs (λmax,em ≈ 550 nm) to 5.5 nm acceptor
However, recent reports of anomalously large polarization QDs (λmax,abs ≈ 590 nm) in homogeneously blended solid-state
effects in QD solids23,24 raise the important question of how films and compare the results to predicted energy transfer rates
energy transfer in close-packed QD solids may differ from that using Förster theory, paying particularly careful attention to the
between QDs and dyes in dilute solution.
Energy transfer within colloidal QD films25,26 is often Received: February 28, 2014
understood in the context of Förster theory, which describes Revised: May 30, 2014
the coupling between two point dipoles. However, the

© XXXX American Chemical Society A dx.doi.org/10.1021/jp502123n | J. Phys. Chem. C XXXX, XXX, XXX−XXX
The Journal of Physical Chemistry C Article

Figure 1. TEM and GISAXS measurements of donor−acceptor separation. (A−C) Short-ligand acceptor. (D−F) Long-ligand acceptor. (A, D) TEM
micrographs of donor−acceptor blended samples, with insets illustrating the corresponding donor−acceptor pair and size histogram revealing the
approximately equimolar bimodal distribution in QD sizes. Scale bar = 40 nm. (B, E) GISAXS patterns for blended 3D films. (C, F) Histograms of
the measured nearest-neighbor interparticle distance (Gaussian fit to TEM data indicated with dotted gray line), demonstrating good agreement
between TEM and GISAXS measurements.

propagation of experimental uncertainty. We obtain simulta- phonic acid), and the other had a shorter aromatic ligand
neous spectral and temporal resolution using spectrally resolved (benzylphosphonic acid). The acceptor QDs were synthesized
transient photoluminescence spectroscopy, where the fluo- identically except for the ligand introduced during overcoating;
rescence lifetime is measured as a function of emission see Dang et al. for further details.6 From TEM images of pure
wavelength. We observe changes in the donor lifetime when close-packed films, the short-ligand acceptor had a diameter of
mixed with a short-ligand acceptor or a long-ligand acceptor 5.5 ± 0.5 nm with mean center-to-center separation of 8 ± 1
and compare to the no-energy-transfer case of a pure donor in nm, while the long-ligand acceptor had a diameter of 5.3 ± 0.5
solution to calculate quenching efficiencies and energy transfer nm with mean center-to-center separation of 9 ± 1 nm.
rates. The energy transfer rates measured from donor Sample Preparation. Molar concentrations of quantum
quenching experiments are more than an order of magnitude dots were estimated using the method of Yu et al., based on the
larger than those predicted, using standard assumptions, by band-edge absorption wavelength and intensity.35 Donor/
Förster theory. We discuss several factors that may be acceptor mixtures were prepared by adding approximately
responsible for this discrepancy, including nonrandom dipole equal moles of each QD species, as determined by absorption
alignment, contributions from higher-order multipoles, and an measurements, and diluting to the desired concentration. QD
enhanced absorption cross section of QDs in close-packed films used for this study were prepared in a nitrogen-filled
solids. These results indicate that care must be taken when glovebox by spin-casting ∼5 mg/mL solutions of QDs in
using Förster theory to make quantitative predictions of energy toluene onto a cleaned glass slide that was previously treated
transfer in colloidal QD solids, which has implications for the with mercaptopropyltrimethoxysilane, resulting in films ap-
design of QD optoelectronic devices such as LEDs and solar proximately 20 nm thick. After spin-coating, the samples were
cells where excitonic energy transfer in the solid state plays a packaged in an inert atmosphere for removal from the glovebox
role in device performance. by placing the QD slide face-down onto a clean glass slide and

■ EXPERIMENTAL METHODS
Materials. CdSe/CdZnS nanocrystals were supplied by QD
sealing the edges with epoxy.
TEM Analysis of Blended Samples. Transmission
electron microscopy (TEM) was used to determine the average
Vision, Inc., and were used as received. QD materials were kept center-to-center distance between donors and acceptors in
under inert atmosphere during all handling and experimenta- blended QD films. Images were taken on a JEOL 2011
tion. QD solutions were prepared with anhydrous solvents operated at 200 kV and using an objective aperture to increase
purchased from Sigma-Aldrich. All QD cores are CdSe with an mass−thickness contrast. Samples for TEM analysis were
overcoated CdZnS shell. Additional details pertaining to these prepared by drop-casting the donor/acceptor mixture onto a
QD materials, including the synthesis procedure, can be found copper TEM grid with amorphous carbon support film. The
in the previous publication by Dang et al.6 TEM analysis donor/acceptor mixtures were diluted such that a monolayer
revealed that the donor QDs used for this study were 4.0 ± 0.5 was formed on the TEM grid, allowing for identification of
nm in diameter, with a center-to-center distance in pure close- individual QDs.
packed monolayers of 6.0 ± 0.9 nm. Two types of acceptor TEM images were processed using ImageJ and then analyzed
QDs were paired with the donor in energy transfer studies: one in Matlab. First, an FFT bandpass filter was applied to the
acceptor had a long-chain aliphatic ligand (octadecylphos- image, which served to better distinguish the edges of the
B dx.doi.org/10.1021/jp502123n | J. Phys. Chem. C XXXX, XXX, XXX−XXX
The Journal of Physical Chemistry C Article

nanocrystals from the carbon support.36 The brightness/ detection system using a calibrated tungsten halogen lamp.
contrast of the image was then maximized, followed by Data were collected from multiple locations on each sample
conversion from grayscale to black and white. Next, particles and averaged together to obtain statistically representative
were indexed and tabulated according to the coordinates of results.
their center of mass, and the area was used to calculate an
effective diameter for each particle. Artifacts of the image
processing algorithm that were clearly not individual QDs were
■ RESULTS AND DISCUSSION
Measurement of Donor−Acceptor Separation. To
rejected based on size or shape. By plotting the histogram of all carefully modulate the donor−acceptor interparticle separation
diameters in an image, two peaks could be resolved which for quantification of energy transfer rates, we first verified that
corresponded to the two different sizes of QDs present in the the length of the ligand bound to the acceptor QD indeed alters
mixture (see insets to Figures 1A,D). We assigned a threshold the donor−acceptor spacing. Approximately equimolar sol-
diameter between these two peaks, creating a population of utions of donor and acceptor QDs were prepared and used
“small” QDs (diameters less than the threshold) and “large” both for TEM and GISAXS analysis as well as fluorescence
QDs (diameters greater than the threshold). Using the centroid lifetime measurements. TEM analysis revealed that the donor
location of each particle in the image, it was then possible to and acceptor species are mostly homogeneously mixed (see the
compute center-to-center distances between every pair of QDs, later section on Experimental Uncertainty for a discussion of
including the case of interest here, which is the distance the effect of incomplete mixing on our Fö rster radius
between “small” (donor) and “large” (acceptor) QDs. The first measurements) and that changing the acceptor ligand from
peak in the histogram of donor−acceptor distances (Figures benzylphosphonic acid to octadecylphosphonic acid does
1C,F) represents the average nearest-neighbor center-to-center indeed alter the interparticle separation (Figure 1; see
distance between each donor and acceptor. The width of this Experimental Methods for a detailed description of the image
peak represents the distribution of different center-to-center processing procedure). A statistical analysis of large-area TEM
distances present in the ensemble. micrographs revealed donor−acceptor distances that were
GISAXS Analysis. The determination of the QD center-to- normally distributed around a mean value. In the short-ligand
center distance by TEM was complemented by grazing- sample, the average donor−acceptor center-to-center distance
incidence small-angle X-ray scattering (GISAXS) analysis of was 7.0 ± 0.9 nm (Figure 1C), whereas in the long-ligand
spin-cast 3D films. GISAXS measurements were performed at sample the distance was 7.8 ± 1.1 nm (Figure 1F).
the undulator-based X9 beamline at the National Synchrotron We note that TEM analysis was performed on monolayer
Light Source at Brookhaven National Laboratory. The X-ray samples, whereas spectroscopic measurements were performed
energy was set to 14.1 keV (wavelength 0.0879 nm), and the on multilayers. Therefore, to verify that the TEM interparticle
beam size was focused to 50 μm tall by 100 μm wide at the separation measurements were representative of the inter-
sample position using a KB mirror system. Because of the particle separation in a 3D film, we measured the grazing-
grazing incidence, beam projection along the beam direction incidence small-angle X-ray scattering (GISAXS) pattern for
was ∼1 cm. The grazing-incident angle of the X-ray beam was samples prepared in the same way as fluorescence lifetime
varied from 0.07° to 0.3°. The presented results are for 0.15°, samples. Figures 1B,E show the GISAXS patterns for the two
which were found to be representative. 2D scattering data were types of mixed films (donor QD blended with each of the two
collected using a Pilatus 1M detector which was calibrated acceptor QDs). From these scattering images, a histogram of
using a silver behenate standard. The QD films used for nearest-neighbor spacing in the 3D blended films was obtained
GISAXS were prepared in a nitrogen-filled glovebox by spin- (see Experimental Methods for detailed procedure). The
casting ∼5 mg/mL solutions of QDs in toluene onto a cleaned GISAXS interparticle distances and the TEM-measured
glass slide that was previously treated with mercaptopropyl- interparticle distances are plotted together in Figures 1C,F.
trimethoxysilane (the same procedure used to prepare samples The measured GISAXS distribution is broader than the TEM
for Förster radius measurements). The interparticle spacing distribution because it includes donor−donor and acceptor−
(Figures 1C,F) was extracted from a given scattering pattern by acceptor distances (in addition to the donor−acceptor distance,
taking a line cut along the in-plane scattering direction (qy) at which is of interest here), whereas our TEM analysis extracts
the Yoneda peak (streak of high intensity just above the sample just the donor−acceptor distance. The excellent agreement
horizon in Figures 1B,E).37 between TEM and GISAXS data inspires confidence in our
Spectrally Resolved Transient Photoluminescence quantification of the donor−acceptor spacing as well as our
Spectroscopy. Transient photoluminescence measurements understanding of its statistical variation within the sample. We
were performed by spectrally resolved time-correlated single- emphasize the importance of obtaining a statistically
photon counting in a modified optical microscope. Briefly, the representative distribution of separation distances before
prepackaged QD samples were loaded onto an inverted performing quantitative analysis of energy transfer data.
microscope with a 60×, 1.4 NA, oil-immersion objective and Characterization methods that isolate just a few donor−
excited at 405 nm with a pulsed diode laser (PicoQuant) acceptor pairs via high-resolution TEM or AFM38 introduce
operating at 5 MHz. Epifluorescence was collected by the observer bias into the measurement and paint an incomplete
objective and reimaged onto the entrance slit of a spectrograph picture of energy transfer.
(Princeton Instruments) where the light was dispersed by a Donor Fluorescence Quenching. To begin the study of
grating. The dispersed fluorescence was either projected onto a energy transfer between donor and acceptor QDs, the first
CCD array (Pixis, Princeton Instruments) for collection of a experimental observation is the quenching of donor fluo-
time-integrated spectrum, or a small wavelength band (∼4 nm) rescence in the presence of the acceptor. For the materials used
was passed through an alternate exit slit for time-resolved here, the emission of the donor overlaps with the second
detection by an avalanche photodiode (τ-SPAD, PicoQuant). absorption feature of both acceptors (Figures 2A,B), indicating
All spectra were corrected for the spectral sensitivity of the that energy transfer from donor to acceptor should be possible.
C dx.doi.org/10.1021/jp502123n | J. Phys. Chem. C XXXX, XXX, XXX−XXX
The Journal of Physical Chemistry C Article

where R0 is the Förster radius (defined to include the donor


quantum efficiency), given by

9ηPLκ 2
6
R0 = 5 4
128π n
∫ λ 4FD(λ)σA(λ) dλ (2)

where ηPL is the quantum efficiency of the donor (0.8 ± 0.1 for
samples used here), κ2 is an orientation factor typically assumed
to be 2/3 for randomly oriented dipoles, n is the refractive
index of the material (typically assumed to be the volume-
weighted sum of the refractive indices of the cores and organic
ligands, equal to about 1.8),25 FD(λ) is the donor emission
spectrum normalized to an integrated value of 1, and σA(λ) is
the acceptor absorption spectrum in units of cross-sectional
area. The absorption cross section for the core−shell QDs used
here was calculated according to the method described by De
Geyter and Hens using core and shell diameters (see expanded
discussion of absorption cross section below).41 From eq 2, the
theoretical Förster radii for the donor/short-ligand-acceptor
Figure 2. (A, B) Emission spectrum of donor (blue) and absorption and donor/long-ligand-acceptor pairs are 5.2 ± 0.2 and 5.1 ±
spectrum of acceptor (red) QDs. The acceptor dots differ only in the 0.2 nm, respectively, resulting in expected energy transfer rates
length of the surface-bound ligand, which is either short (A) or long of 10 (±8) × 106 and 5 (±4) × 106 s−1 per donor−acceptor
(B) as illustrated above each column. (C, D) Emission spectra of the pair. The uncertainty in R0 arises from an estimated 10%
mixed solution (dotted) and film (solid), each normalized to the same uncertainty in both the quantum yield and the overlap integral,
number of counts, of nearly equimolar blends of donor and acceptor while the uncertainty in the associated energy transfer rate
QDs, showing quenching of donor fluorescence in film compared to
stems from the errors in R0, the donor lifetime τ, and (most
solution. The large difference in donor and acceptor emission
intensity, despite similar molar concentrations, results from the significantly) the experimentally measured interparticle separa-
difference in absorption cross section between the large acceptor tion (Figure 1). Based on these calculations, the energy transfer
and small donor QDs at the laser excitation wavelength. rate is expected to differ between the two blends by about a
factor of 2 due to the difference in interparticle separation
imparted by the different ligands.
Experimental Measurement of the Energy Transfer
Indeed, the ratio of donor to acceptor emission intensity Rate. To quantify energy transfer dynamics between donor and
decreased upon deposition from solution to film for both acceptor QDs, we measured the exciton lifetime as a function of
donor−acceptor blends studied here (Figures 2C,D). In emission wavelength (Figure 3). Plots of emission intensity as a
particular, the reduction in relative donor fluorescence is function of emission wavelength and time (Figures 3A,B) show
more significant for donors in the presence of the short-ligand donor quenching and, in particular, reveal more significant
acceptor than the long-ligand, qualitatively confirming initial donor quenching in the case where the donor and acceptor
expectations that ligand length may be used to control energy QDs are closer together (Figure 3A). Spectral and temporal
transfer efficiency between donors and acceptors in otherwise resolution allows for clear distinction between donor and
identical materials. These observations suggesting the occur- acceptor fluorescence. To obtain population decay curves
rence of energy transfer are further strengthened and quantified (panels C−F of Figure 3), the partially overlapping emission
by measurements of changes in the donor lifetime in the spectra from donor and acceptor QDs were separated by fitting
presence versus absence of acceptors (described in detail two Gaussians to the emission spectrum recorded at each time
below). Changes in donor lifetime conclusively indicate near- slice of the data. The intensities in the donor components of
field energy transfer and rule out other possible mechanisms to the double Gaussian fit were then spectrally integrated to
explain changes in the donor intensity, such as emission/ calculate a donor lifetime in the presence of acceptors. The
reabsorption. In the following sections, we first calculate the exciton population decay of the donor in the presence of the
expected energy transfer rates using the overlap integral and the acceptor in solid films is plotted in Figure 3, panels C−F, and
measured distance between donors and acceptors and then compared to the donor lifetime in solution where no energy
compare these theoretical predictions to the experimentally transfer occurs. The dramatic change in lifetime results from
measured energy transfer rates based on changes in the donor energy transfer to neighboring acceptors, and the per-pair
lifetime in the absence or presence of acceptors. energy transfer rate may be extracted from the lifetime change
Calculation of Expected Energy Transfer Dynamics as described below in order to compare to theoretical
Using Fö rster Theory. Theoretical energy transfer rates may predictions presented in the previous section. The presence
be calculated using the physical properties of the donor− of the QD acceptors in solid film reduces donor fluorescence
acceptor blend.39,40 In this case, the energy transfer rate (kET) is lifetime as well as intensity, indicating near-field energy transfer
inversely proportional to the center-to-center distance between between donor and acceptor QDs.
QDs (d) and the total lifetime of the donor (τD) The change in donor lifetime observed by spectrally resolved
transient photoluminescence (Figure 3) was used to calculate
1 ⎛ R0 ⎞
6
kET = ⎜ ⎟ the energy transfer rate between donor−acceptor pairs in the
τD ⎝ d ⎠ (1) blended solid films. In the absence of energy transfer, the total
D dx.doi.org/10.1021/jp502123n | J. Phys. Chem. C XXXX, XXX, XXX−XXX
The Journal of Physical Chemistry C Article

which can be used, in turn, to calculate the effective Förster


radius.
Correctly applying eq 5 requires comparison between the
donor lifetime in the absence and presence of energy transfer.
For this purpose, it is essential to use the solution decay rate for
kpure rather than the pure film lifetime because energy transfer
within the inhomogeneously broadened pure donor ensemble26
and exciton diffusion to quenching sites3,43 obfuscates the true
intrinsic decay rate of a QD in the solid film. We verify that
energy transfer occurs within the pure donor film by
observation of a spectral red-shift over the course of the
exciton lifetime (Figures 4B,D), resulting in an emission-

Figure 3. Spectrally resolved transient photoluminescence measure-


ments of donor−acceptor blended films. The left and right columns
correspond to the short-ligand and long-ligand acceptor, respectively.
(A, B) Normalized fluorescence intensity, indicated by the color bar, Figure 4. Spectrally resolved transient photoluminescence of (A) pure
plotted as a function of time and emission wavelength. Spectrally donor solution and (B) pure donor film. Population decay at selected
integrated donor counts are plotted on both linear (C, D) and wavelengths are shown in panels C and D for the donor solution and
semilogarithmic (E, F) axes. Solid black lines are fits to the film, respectively. Decay traces range in wavelength from 520 nm
experimental data (open circles). (purple) to 590 nm (red), in increments of 10 nm.

fluorescence decay rate (kpure) is the sum of the intrinsic


radiative (kr) and nonradiative (knr) decay rates
energy-dependent lifetime not observed in donor solutions
k pure = k r + k nr (3) (Figures 4A,C).26 We note that the lifetime of the reddest QDs
in the pure donor film approaches the lifetime of the donor
In the mixed donor−acceptor film, the total fluorescence decay QDs in solution. This observation further confirms that the
rate (kmix) is solution lifetime is the inherent no-energy-transfer lifetime of
k mix = k r + k nr + ∑ kET the QDs in the ensemble, as these reddest dots have no
(4)
neighbors capable of accepting a band-edge exciton and are
where the summation is over all possible FRET acceptors. therefore effectively isolated. Noninteracting QDs in solution
Therefore, the total energy transfer rate may be calculated by and film have the same inherent lifetime, so the solution lifetime
subtracting eq 3 from eq 4. Because of the 1/d6 distance scaling, and quantum efficiency (rather than the pure donor film
only nearest neighbors are expected to contribute substantially lifetime and quantum efficiency) must be used to predict
to the total energy transfer rate.26 Therefore, the per-pair energy transfer dynamics in blended QD assemblies.
energy transfer rate (kET,pp) can be calculated from In the blended QD films, the distribution of donor−acceptor
1 distances (Figure 1) is expected to result in a distribution of
kET,pp = (k mix − k pure) energy transfer rates and, accordingly, a distribution of donor
N (5)
fluorescence lifetimes. To obtain a statistically averaged lifetime,
where N is the number of nearest neighbors. Danisch et al. we first fit the donor fluorescence decay to a multiexponential
previously determined the approximate number of nearest function to deconvolve the time-dependent exciton population
neighbors of a given size, given a random close packed lattice of n(t) from the instrument response function (determined almost
two different sized spheres with size ratio 1:1.4 (the ratio of entirely by the temporal width of the laser pulse). The
donor and acceptor QD diameters in our study).42 For an statistically averaged lifetime, τavg, is then obtained via
equimolar mixture of donor and acceptor QDs with this size ∞
ratio, N equals 3 (see expanded discussion of experimental ∫0 tn(t ) dt
uncertainty below). Using eq 5, the lifetime comparison yields a τavg = ∞
direct experimental measurement of the energy transfer rate,
∫0 n(t ) dt (6)

E dx.doi.org/10.1021/jp502123n | J. Phys. Chem. C XXXX, XXX, XXX−XXX


The Journal of Physical Chemistry C Article

Table 1. Summary of Theoretical and Experimental Energy Transfer Rates (per Donor−Acceptor pair, pp)
measured donor−acceptor distance kET,pp (s−1) from
(nm) R0 (nm) from theory kET,pp (s−1) from theory experiment R0 (nm) from experiment
short ligand 7.0 (±0.9) 5.2 (±0.2) 10 (±8) × 106 260 (±60) × 106 9 (±2)
long ligand 7.8 (±1.1) 5.1 (±0.2) 5 (±4) × 106 60 (±10) × 106 8 (±1)

The lifetime of the donor in solution is 15.5 ± 0.3 ns (kpure = films are only ∼3 QD layers thick and surface QDs have fewer
0.065 ± 0.001 ns−1), while the average lifetime of the donor in nearest neighbors, we have probably underestimated the per-pair
the short-ligand acceptor blend is 1.2 ± 0.2 ns (kmix = 0.8 ± 0.1 energy transfer rate by choosing a bulk value for N.
ns−1) and the average lifetime of the donor in the long-ligand Additionally, the TEM images shown in Figure 1 indicate
acceptor blend is 4.1 ± 0.2 ns (kmix = 0.24 ± 0.01 ns−1) (Figure that the QDs are not entirely randomly distributed in a
3). Using these values and eq 5, we find that the energy transfer homogeneously blended film. If slight aggregation of similar
rate per donor−acceptor pair in the short-ligand blend is kET,pp types of QDs occurs within our 3D films, this would also
= 2.6 (±0.6) × 108 s−1, while for the long-ligand blend the decrease the true value of N to less than 3. A smaller number of
energy transfer rate is kET,pp = 0.6 (±0.1) × 108 s−1. An nearest acceptor neighbors per donor QD further exaggerates
experimental R0 can be calculated from these values by the discrepancy between theoretical and measured energy
rearranging eq 1 transfer rates and reinforces our claim that actual energy
transfer rates are faster than those predicted by standard theory.
R 0 = d(τpurekET,pp)1/6 (7) These factors, as well as errors in measurement of quantum
efficiency, lifetime, and other experimental parameters, have
The experimentally determined values of R0 are 9 ± 2 nm for been quantitatively accounted for in the propagated uncertainty
the donor-short-ligand acceptor pair and 8 ± 1 nm for the and are insufficient in explaining the difference between
donor-long ligand acceptor pair. These numbers are expected and measured energy transfer rates. We believe that
summarized in Table 1. other factors play a role beyond those usually considered in the
Discrepancy between Theory and Experiment. The calculation of the Förster radius. We consider some possibilities
theoretical Förster radii and energy transfer rates calculated
below.
from materials properties are substantially smaller than those
Dipole Orientation. As seen in the TEM images of Figure 1,
measured experimentally (see Table 1). The expected
the QDs used in this study are not spherical but instead have a
theoretical per-pair energy transfer rates based on the overlap
nearly tetrahedral geometry, as discussed in a previous study
integral and interparticle separation are 10 (±8) × 106 and 5
that utilized these same materials.6 This may lead to some
(±4) × 106 s−1 for the donor paired with the short-ligand
oriented packing and alignment of the QDs in thin film,
acceptor and long-ligand acceptor, respectively. However, the
measured energy transfer rates are more than an order of resulting in a transition dipole that is not random but instead
magnitude larger, 260 (±60) × 106 and 60 (±10) × 106 s−1 for aligned. Furthermore, the observed 2D character of the
the two cases, respectively (Table 1). Calculated Förster energy emission transition dipole in CdSe QDs makes better-than-
transfer rates to second nearest neighbors are on the order of random alignment between donor emission and acceptor
104 s−1 and therefore contribute negligibly to the total per-pair absorption transition dipole moments even more plausible.44
energy transfer rate. The discrepancy between theoretical Murray et al. found that QDs within superlattices tend to align
predictions and experimental measurements in our model their crystal axes,45 suggesting ready alignment of transition
donor−acceptor system suggests that application of Förster’s dipoles in some QD assemblies. If the transition dipoles were
equation to solid state QD systems requires further perfectly aligned (κ2 = 4),27 this would increase the calculated
consideration, as we now address. R0 from 5.1 nm to about 7.0 nm and thereby increase the
Experimental Uncertainty. Obvious experimental uncer- theoretical energy transfer rate by about a factor of 7 for both
tainty cannot account for the magnitude of the difference films. This cannot completely account for the discrepancy
between experimentally measured and theoretically predicted between theory and experiment, and furthermore, it is unlikely
energy transfer rates and Förster radii. There is some systematic that all of the transition dipoles in our mixed-size films are
error in the measurement of donor−acceptor separation perfectly aligned because of geometric packing restrictions.
distance due to an inability of the image processing algorithm Consequently, dipole orientation is likely a contributing cause
to distinguish QDs that appear very close together in the TEM of the observed discrepancy, but only a minority factor.
micrograph. This leads to a systematic overestimation of the QD Shape. In addition to the QD shape affecting packing
interparticle separation on the order of a few tenths of and transition dipole alignment, it is also possible that shape
nanometers, but this error is small compared to the propagated affects Coulombic and dipole−dipole coupling between
uncertainty of ∼1.0 nm for all calculations and cannot explain neighboring nanocrystals.46 Several reports in the literature
the discrepancy between theory and experiment. Moreover, the have demonstrated different energy transfer behavior in systems
excellent agreement between TEM and GISAXS data suggests of QDs compared to quantum rods or dot-in-rod struc-
that the propagated experimental uncertainty captures any error tures.46,47 On the basis of published reports of the magnitude of
in interparticle distance. this affect, we estimate that shape could account for up to 10−
Experimental determination of kET and R0 depends on 20% of the discrepancy.
knowledge of the number of nearest neighbors, as specified in Higher Order Multipoles. In addition to these physical
eq 5. However, even if there are twice as many nearest acceptor factors that may be incorrectly accounted for in the usual
neighbors as we expect (N = 6 instead of N = 3), the calculations of R 0 and k ET , the discrepancy may be
experimentally measured energy transfer rate remains ∼5−15 fundamentally due to the assumption that QD−QD energy
times larger than the predicted value. Further, because our QD transfer in thin film is mediated by point dipole−point dipole
F dx.doi.org/10.1021/jp502123n | J. Phys. Chem. C XXXX, XXX, XXX−XXX
The Journal of Physical Chemistry C Article

coupling. Baer and Rabani showed that for small surface-to- absorption enhancement in the denser QD array. These factors
surface separations dipole−quadrupole coupling may contrib- suggest that the results of a full ligand series to incrementally
ute non-negligibly to the total energy transfer rate.48 In change the interparticle separation may be difficult to interpret.
particular, for surface-to-surface distances less than about 20% Even if Förster theory is the correct theoretical framework, such
of the total center-to-center distance, dipole−quadrupole a study would likely not result in the expected 1/d6 scaling
coupling may contribute more than a third to the combined because the polarizability (acceptor absorptivity) of the
energy transfer rate. For this model system, the expected nanocrystal assembly, its refractive index, and the interparticle
difference in energy transfer rates from donor to acceptor for separation all change simultaneously in different directions.
the two test cases was a factor of 2, as described above;
however, a factor of 4 difference was observed experimentally.
The relative change in energy transfer rate from the donor to
■ CONCLUSIONS
We have measured the energy transfer rate between donor and
the short-ligand acceptor vs long-ligand acceptor may be due in acceptor QDs in homogeneously blended films, where the
part to dipole−multipole interactions as the surface-to-surface interparticle separation was tuned through altering the ligand
separation is 25% of the total center-to-center separation in the length of the acceptor. The measured energy transfer rates were
case of the short-ligand acceptor (putting it into the realm of found to be more than an order of magnitude larger than the
more significant dipole−quadrupole coupling) but more than theoretical energy transfer rates calculated based on Förster
40% of the center-to-center distance in the case of the long theory, using standard assumptions. Several possible sources of
ligand acceptor. As a result, the larger than expected difference the discrepancy were identified and quantified, including QD
between the measured energy transfer rates with the two shape, preferential orientation of transition dipoles, enhanced
acceptors could be due in part to dipole−quadrupole coupling, absorption cross section, and dipole−multipole coupling; these
which contributes to the energy transfer rate in the case of the factors may result in different behavior for QDs in solid films
shorter-ligand acceptor but has a negligible effect for the longer- than for QDs in solution. To further elucidate the mechanisms
ligand acceptor. Contributions from dipole−quadrupole underlying energy transfer in CdSe QD assemblies, further
coupling may account for part of the magnitude difference experimentation may focus on the direct comparison of energy
between the measured and theoretical energy transfer rates, but transfer between isolated donor−acceptor pairs in solution to
the exact contribution is difficult to quantify. that in solid film. This work suggests that predicting energy
Absorption Cross Section. The absorption cross section of transfer between QDs at small distances within a close-packed
the acceptor QDs sensitively affects the calculated overlap film requires careful consideration and informs our under-
integral, and therefore the Förster radius, for a donor−acceptor standing of excitonic energy transfer in QD films relevant for
pair. However, there are inconsistencies in the literature as to optoelectronic devices such as solar cells and LEDs.
the magnitude of the absorption cross section as a function of
QD size. In this work, the calculated Förster radius changes
between 4.8 and 6.2 nm depending on the value of the
■ AUTHOR INFORMATION
Corresponding Author
absorption cross section used. CdZnS shells surround the *E-mail tisdale@mit.edu (W.A.T.).
nanoparticles used in this study, which complicates the
determination of absorption cross section. As a result, the Notes
The authors declare no competing financial interest.


work of De Geyter et al., which directly addresses this issue, was
used for calculating absorption cross section.41 However, other
methods35,49,50 for calculating the absorption cross section can ACKNOWLEDGMENTS
yield slightly different values that may change the absorption The authors thank Liza Lee for helpful discussions, Kevin Yager
cross section and therefore the Förster radius. The lack of for performing the GISAXS measurements, and Jonathan
experimental consensus renders exact calculation of the Steckel and Seth Coe-Sullivan of QD Vision, Inc., for supplying
absorption cross section difficult and highlights the need for QD materials. This work was supported as part of the Center
improved data in this important area of QD research. for Excitonics, an Energy Frontier Research Center funded by
In addition to variation in the calculated absorption cross the U.S. Department of Energy, Office of Science, Office of
section depending on method, a further likely contribution to Basic Energy Sciences, under Award DE-SC0001088 (MIT).
the larger than expected energy transfer rate is an enhanced A.J.M. and M.C.W. gratefully acknowledge fellowship support
absorption cross section of the acceptor QDs in film compared from the National Science Foundation Graduate Research
to solution. Recently, Geiregat et al. showed that the absorption Fellowship Program under Grant No. 1122374. GISAXS
cross section per QD in close-packed films is enhanced relative measurements were performed at the Center for Functional
to the solution value due to dipole-induced-dipole effects in the Nanomaterials and the National Synchrotron Light Source,
solid.23 For CdSe QDs of the sizes used here, the absorption Brookhaven National Laboratory, which is supported by the
cross section in thin film was measured to be 3−4 times larger U.S. Department of Energy, Office of Basic Energy Sciences,
than the absorption cross section in solution.23 If the measured under Contract DE-AC02-98CH10886.
absorption spectrum is scaled to this new, larger magnitude
value (and possible dipole orientation effects ignored), the
theoretical value for R0 increases from 5.1−5.2 to 6.2 nm for
■ ABBREVIATIONS
QD, quantum dot; FRET, Förster resonant energy transfer;
both donor−acceptor pairs. Furthermore, these polarization CdSe, cadmium selenide; CdZnS, cadmium zinc sulfide.
effects should play a larger role in short-ligand QD assemblies
due to closer proximity of neighboring QDs.23 We indeed
observe a larger deviation from the predicted theoretical energy
■ REFERENCES
(1) Colvin, V. L.; Schlamp, M. C.; Alivisatos, A. P. Light-Emitting
transfer rate in the system with the shorter ligand acceptor than Diodes Made from Cadmium Selenide Nanocrystals and a Semi-
that with the longer ligand acceptor, consistent with stronger conducting Polymer. Nature 1994, 370, 354−357.

G dx.doi.org/10.1021/jp502123n | J. Phys. Chem. C XXXX, XXX, XXX−XXX


The Journal of Physical Chemistry C Article

(2) Coe, S.; Woo, W. K.; Bawendi, M.; Bulovic, V. Electro- Semiconductor Quantum Dot Photoluminescence by Proximal Gold
luminescence from Single Monolayers of Nanocrystals in Molecular Nanoparticles. Nano Lett. 2007, 7, 3157−3164.
Organic Devices. Nature 2002, 420, 800−803. (23) Geiregat, P.; Justo, Y.; Abe, S.; Flamee, S.; Hens, Z. Giant and
(3) Shirasaki, Y.; Supran, G. J.; Bawendi, M. G.; Bulovic, V. Broad-Band Absorption Enhancement in Colloidal Quantum Dot
Emergence of Colloidal Quantum-Dot Light-Emitting Technologies. Monolayers through Dipolar Coupling. ACS Nano 2013, 7, 987−993.
Nat. Photonics 2013, 7, 13−23. (24) Wolcott, A.; Doyeux, V.; Nelson, C. A.; Gearba, R.; Lei, K. W.;
(4) Kamat, P. V. Quantum Dot Solar Cells. Semiconductor Yager, K. G.; Dolocan, A. D.; Williams, K.; Daniel, N.; Zhu, X. Y.
Nanocrystals as Light Harvesters. J. Phys. Chem. C 2008, 112, Anomalously Large Polarization Effect Responsible for Excitonic Red-
18737−18753. Shifts in PbSe Quantum Dot Solids. J. Phys. Chem. Lett. 2011, 2, 795−
(5) Pattantyus-Abraham, A. G.; Kramer, I. J.; Barkhouse, A. R.; Wang, 800.
X.; Konstantatos, G.; Debnath, R.; Levina, L.; Raabe, I.; Nazeeruddin, (25) Kagan, C.; Murray, C.; Nirmal, M.; Bawendi, M. Electronic
M. K.; Sargent, E. H. Depleted-Heterojunction Colloidal Quantum Energy Transfer in CdSe Quantum Dot Solids. Phys. Rev. Lett. 1996,
Dot Solar Cells. ACS Nano 2010, 4, 3374−3380. 76, 1517−1520.
(6) Dang, C.; Lee, J.; Breen, C.; Steckel, J. S.; Coe-Sullivan, S.; (26) Crooker, S.; Hollingsworth, J.; Tretiak, S.; Klimov, V. Spectrally
Nurmikko, A. Red, Green and Blue Lasing Enabled by Single-Exciton Resolved Dynamics of Energy Transfer in Quantum-Dot Assemblies:
Gain in Colloidal Quantum Dot Films. Nat. Nanotechnol. 2012, 335− Towards Engineered Energy Flows in Artificial Materials. Phys. Rev.
339. Lett. 2002, 89, 186802.
(7) Klimov, V. I.; Ivanov, S. A.; Nanda, J.; Achermann, M.; Bezel, I.; (27) Medintz, I.; Hildebrandt, N. FRET - Förster Resonance Energy
McGuire, J. A.; Piryatinski, A. Single-Exciton Optical Gain in Transfer: From Theory to Applications; Wiley-VCH: Weinheim,
Semiconductor Nanocrystals. Nature 2007, 447, 441−446. Germany, 2013.
(8) Konstantatos, G.; Howard, I.; Fischer, A.; Hoogland, S.; Clifford, (28) Kagan, C.; Murray, C.; Bawendi, M. Long-Range Resonance
J.; Klem, E.; Levina, L.; Sargent, E. H. Ultrasensitive Solution-Cast Transfer of Electronic Excitations in Close-Packed CdSe Quantum-
Quantum Dot Photodetectors. Nature 2006, 442, 180−183. Dot Solids. Phys. Rev. B 1996, 54, 8633−8643.
(9) Shcherbatyuk, G. V.; Inman, R. H.; Wang, C.; Winston, R.; (29) Lunz, M.; Bradley, A. L.; Chen, W.-y.; Gun’ko, Y. K. Two-
Ghosh, S. Viability of Using Near Infrared PbS Quantum Dots as Dimensional Forster Resonant Energy Transfer in a Mixed Quantum
Active Materials in Luminescent Solar Concentrators. Appl. Phys. Lett. Dot Monolayer: Experiment and Theory. J. Phys. Chem. C 2009, 113,
2010, 96, 191901. 3084−3088.
(10) Bruchez, M.; Moronne, M.; Gin, P.; Weiss, S.; Alivisatos, A. P. (30) Kim, D.; Okahara, S.; Nakayama, M.; Shim, Y. Experimental
Semiconductor Nanocrystals as Fluorescent Biological Labels. Science Verification of Förster Energy Transfer between Semiconductor
Quantum Dots. Phys. Rev. B 2008, 78, 153301.
1998, 281, 2013−2016.
(31) Moreels, I.; Justo, Y.; Rainò, G.; Stöferle, T.; Hens, Z.; Mahrt, R.
(11) Medintz, I. L.; Uyeda, H. T.; Goldman, E. R.; Mattoussi, H.
F. Impact of the Band-Edge Fine Structure on the Energy Transfer
Quantum Dot Bioconjugates for Imaging, Labelling and Sensing. Nat.
between Colloidal Quantum Dots. Adv. Opt. Mater. 2013, 2, 126−130.
Mater. 2005, 4, 435−46.
(32) Lingley, Z.; Lu, S.; Madhukar, A. A High Quantum Efficiency
(12) Bourzac, K. Quantum Dots Go on Display. Nature 2013, 493,
Preserving Approach to Ligand Exchange on Lead Sulfide Quantum
283−283.
Dots and Interdot Resonant Energy Transfer. Nano Lett. 2011, 11,
(13) Anikeeva, P. O.; Halpert, J. E.; Bawendi, M. G.; Bulović, V.
2887−2891.
Electroluminescence from a Mixed Red-Green-Blue Colloidal
(33) Lunz, M.; Bradley, A. L.; Chen, W.-Y.; Gerard, V. A.; Byrne, S.
Quantum Dot Monolayer. Nano Lett. 2007, 7, 2196−200.
J.; Gun’ko, Y. K.; Lesnyak, V.; Gaponik, N. Influence of Quantum Dot
(14) Akselrod, G. M.; Prins, F.; Poulikakos, L. V.; Lee, E. M. Y.;
Concentration on Förster Resonant Energy Transfer in Monodis-
Weidman, M. C.; Mork, A. J.; Willard, A. P.; Bulović, V.; Tisdale, W. A. persed Nanocrystal Quantum Dot Monolayers. Phys. Rev. B 2010, 81,
Subdiffusive Exciton Transport in Quantum Dot Solids. Nano Lett. 205316.
2014, in press. (34) Poulikakos, L. V.; Prins, F.; Tisdale, W. A. Transition from
(15) Achermann, M.; Petruska, M. A.; Crooker, S. A.; Klimov, V. I. Thermodynamic to Kinetic-Limited Excitonic Energy Migration in
Picosecond Energy Transfer in Quantum Dot Langmuir-Blodgett Colloidal Quantum Dot Solids. J. Phys. Chem. C 2014, 118, 7894−
Nanoassemblies. J. Phys. Chem. B 2003, 107, 13782−13787. 7900.
(16) Xu, F.; Ma, X.; Haughn, C. R.; Benavides, J.; Doty, M. F.; (35) Yu, W. W.; Qu, L.; Guo, W.; Peng, X. Experimental
Cloutier, S. G. Efficient Exciton Funneling in Cascaded PbS Quantum Determination of the Extinction Coefficient of CdTe, CdSe, and
Dot Superstructures. ACS Nano 2011, 5, 9950−9957. CdS Nanocrystals. Chem. Mater. 2003, 125, 2854−2860.
(17) Medintz, I. L.; Clapp, A. R.; Brunel, F. M.; Tiefenbrunn, T.; (36) Segets, D.; Lucas, J. M.; Klupp Taylor, R. N.; Scheele, M.;
Uyeda, H. T.; Chang, E. L.; Deschamps, J. R.; Dawson, P. E.; Zheng, H.; Alivisatos, A. P.; Peukert, W. Determination of the
Mattoussi, H. Proteolytic Activity Monitored by Fluorescence Quantum Dot Band Gap Dependence on Particle Size from Optical
Resonance Energy Transfer through Quantum-Dot-Peptide Con- Absorbance and Transmission Electron Microscopy Measurements.
jugates. Nat. Mater. 2006, 5, 581−9. ACS Nano 2012, 6, 9021−9032.
(18) Medintz, I. L.; Clapp, A. R.; Mattoussi, H.; Goldman, E. R.; (37) Yoneda, Y. Anomalous Surface Reflection of X Rays. Phys. Rev.
Fisher, B.; Mauro, J. M. Self-Assembled Nanoscale Biosensors Based 1963, 131, 2010−2013.
on Quantum Dot FRET Donors. Nat. Mater. 2003, 2, 630−8. (38) Zheng, K.; Ž ídek, K.; Abdellah, M.; Zhu, N.; Chábera, P.;
(19) Allan, G.; Delerue, C. Energy Transfer between Semiconductor Lenngren, N.; Chi, Q.; Pullerits, T. Directed Energy Transfer in Films
Nanocrystals: Validity of Förster’s Theory. Phys. Rev. B 2007, 75, of CdSe Quantum Dots: Beyond the Point Dipole Approximation. J.
195311. Am. Chem. Soc. 2014, 136, 6259−6268.
(20) Curutchet, C.; Franceschetti, A.; Zunger, A.; Scholes, G. D. (39) Fö rster, T. Zwischenmolekulare Energiewanderung Und
Examining Forster Energy Transfer for Semiconductor Nanocrystalline Fluoreszenz. Ann. Phys. 1948, 437, 55−75.
Quantum Dot Donors and Acceptors. J. Phys. Chem. C 2008, 112, (40) Powell, R. C.; Soos, Z. G. Singlet Exciton Energy Transfer in
13336−13341. Organic Solids. J. Lumin. 1975, 11, 1−45.
(21) Medintz, I. L.; Mattoussi, H. Quantum Dot-Based Resonance (41) De Geyter, B.; Hens, Z. The Absorption Coefficient of PbSe/
Energy Transfer and Its Growing Application in Biology. Phys. Chem. CdSe Core/Shell Colloidal Quantum Dots. Appl. Phys. Lett. 2010, 97,
Chem. Phys. 2009, 11, 17−45. 161908.
(22) Pons, T.; Medintz, I. L.; Sapsford, K. E.; Higashiya, S.; Grimes, (42) Danisch, M.; Jin, Y.; Makse, H. A. Model of Random Packings
A. F.; English, D. S.; Mattoussi, H. On the Quenching of of Different Size Balls. Phys. Rev. E 2010, 81, 051303.

H dx.doi.org/10.1021/jp502123n | J. Phys. Chem. C XXXX, XXX, XXX−XXX


The Journal of Physical Chemistry C Article

(43) Miyazaki, J.; Kinoshita, S. Site-Selective Spectroscopic Study on


the Dynamics of Exciton Hopping in an Array of Inhomogeneously
Broadened Quantum Dots. Phys. Rev. B 2012, 86.
(44) Chung, I.; Shimizu, K. T.; Bawendi, M. G. Room Temperature
Measurements of the 3D Orientation of Single CdSe Quantum Dots
Using Polarization Microscopy. Proc. Natl. Acad. Sci. U. S. A. 2003,
100, 405−408.
(45) Murray, C. B.; Kagan, C. R.; Bawendi, M. G. Self-Organization
of CdSe Nanocrystallites into Three-Dimensional Quantum Dot
Superlattices. Science 1995, 270, 1335−1338.
(46) Schrier, J.; Wang, L.-W. Shape Dependence of Resonant Energy
Transfer between Semiconductor Nanocrystals. J. Phys. Chem. C 2008,
112, 11158−11161.
(47) Halivni, S.; Sitt, A.; Hadar, I.; Banin, U. Effect of Nanoparticle
Dimensionality on Fluorescence Resonance Energy Transfer in
Nanoparticle−Dye Conjugated Systems. ACS Nano 2012, 6, 2758−
2765.
(48) Baer, R.; Rabani, E. Theory of Resonance Energy Transfer
Involving Nanocrystals: The Role of High Multipoles. J. Chem. Phys.
2008, 128, 184710.
(49) Jasieniak, J.; Smith, L.; Embden, J. V.; Mulvaney, P.; Califano,
M. Re-Examination of the Size-Dependent Absorption Properties of
CdSe Quantum Dots. J. Phys. Chem. C 2009, 113, 19468−19474.
(50) Leatherdale, C. A.; Mikulec, F. V.; Bawendi, M. G. On the
Absorption Cross Section of CdSe Nanocrystal Quantum Dots. J. Phys.
Chem. B 2002, 106, 7619−7622.

I dx.doi.org/10.1021/jp502123n | J. Phys. Chem. C XXXX, XXX, XXX−XXX

You might also like