Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

Letter

pubs.acs.org/JPCL

Ultrafast Charge Separation from Highly Reductive ZnTe/CdSe Type


II Quantum Dots
Shengye Jin,†,‡,∥ Jun Zhang,†,∥ Richard D. Schaller,†,§ Tijana Rajh,*,† and Gary P. Wiederrecht*,†,‡

Center for Nanoscale Materials, Argonne National Laboratory, Argonne, Illinois 60439, United States

Argonne−Northwestern Solar Energy Research Center, Northwestern University, Evanston, Illinois 60208, United States
§
Department of Chemistry, Northwestern University, Evanston, Illinois 60208, United States
*
S Supporting Information

ABSTRACT: The low electron affinity of ZnTe quantum dots (QDs)


makes it of interest for critically important photocatalytic processes.
However, studies of charge transfer from such materials to adsorbates
have not been reported. Here, the ultrafast internal and external charge
separation dynamics in ZnTe/CdSe core/shell type II QDs have been
studied through femtosecond transient absorption spectroscopy. The
internal electron transfer time from a ZnTe core to a CdSe shell was
found to be 0.67 ps, while the external electron transfer time from QDs
to adsorbed molecules was found to be <0.2 ps. Such a fast external
charge separation time is due to the extraordinarily high conduction
band energy potential of this QD composition. This study indicates the
preservation of high photodriven reductive abilities in the ZnTe/CdSe
core/shell type II heterostructures.
SECTION: Spectroscopy, Photochemistry, and Excited States

C ore/shell type II semiconductor quantum dots (QDs) are


of great interest for photovoltaic and photocatalytic
applications because the edge-state positions of the core and
driven reductive potentials than other type II heterostructures.
The core/shell structure also acts to protect the ZnTe core
from irreversible photooxidation damage while largely main-
shell allow for the separation of photon-induced electron hole taining its highly reductive properties.
pairs across the core−shell interface.1−10 This is in sharp Most current studies on ZnTe materials primarily focus on
contrast to type I QDs, where both the electron and hole are synthesis.14−16 A few studies report the photophysical proper-
confined in either the core or shell. Type II QDs possess many ties of type II heterostructures containing ZnTe as the shell for
unique properties, such as long charge carrier lifetimes,3,7 fast hole trapping.17−19 However, the photoinduced internal and
charge separation dynamics,10,11 and red-extended absorption external charge separation and recombination dynamics in type
spectra.10 Among the various semiconductor materials used to II heterostructures with ZnTe as the core are still unknown, but
fabricate type II QDs, ZnTe draws particular interest because of crucial for utilizing ZnTe materials in solar energy conversion.
its low electron affinity (high conduction band (CB) energy In this Letter, we report the studies of internal and external
level: −2.7 eV vs Vac. in bulk with a 2.3 eV band gap).12,13 Such charge separation and recombination dynamics in a ZnTe/
a high CB offers substantial benefits when forming a type II CdSe core/shell type II QD as well as a QD−AQ
heterostructure by offering a large energy range over which to (anthraquinone-2,3-dicarboxylic acid) complex as shown in
select different core or shell materials. The high CB also allows
Figure 1b. Following photoexcitation, an internal electron
for additional freedom to tune the degree of quantum
transfer (ET) process (kIET) can occur from the ZnTe core to
confinement by changing core size and shell thickness to
the CdSe shell because of the type II energy alignment, which is
achieve efficient electron and hole separation. Comparison of
the examples of CdTe/CdSe and ZnTe/CdSe shown in Figure followed by an external ET (kEET) from the CdSe shell to AQs.
1a illustrate this point. In a CdTe/CdSe core/shell type II QD, The separated electron and hole can recombine through
the limited energy difference between CBs in these two internal (kIR) or external (kER) charge recombination processes.
materials requires a large CdSe shell thickness to maintain a Through transient absorption (TA) studies we found that the
type II band alignment.5 However, in a ZnTe/CdSe core/shell internal ET time is ∼0.67 ps, and the external ET time is <0.2
QD, the significantly higher CB energy level of ZnTe allows a ps, which is ∼10-fold faster than that in CdTe/CdSe type II
much thinner CdSe shell to still achieve efficient separation of
the electron and hole. A thinner shell thickness means stronger Received: July 5, 2012
quantum confinement and hence higher electron energy levels Accepted: July 19, 2012
in it. Therefore, ZnTe/CdSe QDs can afford higher photo- Published: July 23, 2012

© 2012 American Chemical Society 2052 dx.doi.org/10.1021/jz3008886 | J. Phys. Chem. Lett. 2012, 3, 2052−2058
The Journal of Physical Chemistry Letters Letter

Figure 2. UV−vis absorption (solid) and emission (dashed) spectra of


ZnTe/CdSe core/shell type II QDs, as well as the absorption of the
ZnTe seeds (no emission is observed in ZnTe seeds). The energy level
diagram of the QD is shown in the inset. Solid blocks indicate bulk CB
and valence band (VB) energy levels. The arrows indicate three
possible electronic transitions (T1, T2, and T3), which are also
marked on the corresponding absorption bands. T1 is assigned to the
transition from the lowest energy valence-band 1sh level in ZnTe
(−5.1 eV) to a delocalized conduction-band electron level in ZnTe
(−2.6 eV). T2 is assigned to the spatially indirect transition from the
valence-band 1sh in ZnTe (−5.1 eV) to the lowest energy conduction-
band 1se level in CdSe (−3.1 eV). T3 is assigned to the transition from
the lowest energy delocalized valence-band hole level (−5.6 eV) to the
conduction-band 1se level in CdSe (−3.1 eV).

lifetime component with a time constant of ∼40 ns. Such a long


exciton lifetime is not observed in the TA kinetics of a ZnTe
seed QD (Figure S5), and is strong evidence of the spatial
separation of the electron in CdSe and the hole in ZnTe.11,20,21
The fast component with a lifetime of 0.25 ns in the
Figure 1. (a) Energy level diagrams of ZnTe/CdSe and CdTe/CdSe fluorescence decay may be partially due to the electron and/
core/shell type II QDs. Solid blocks indicate bulk CB and VB energy or hole transfer to the defect states in and around the QDs.
levels. Dashed lines are the estimated lowest energy electron and hole
levels considering quantum confinement. (b) The photoinduced The TA spectra of ZnTe/CdSe QDs at different delay times
internal and external ET and recombination processes in ZnTe/ after excitation at 400 nm are shown in Figure 3a. Consistent
CdSe−AQ complexes. with its UV−vis absorption spectrum, clear bleach bands
centered at 500 and 600 nm are observed. The bleaches are due
to the state-filling of the electron in the CB and the hole in the
QDs.10 The findings in this study reveal efficient photophysical VB levels associated with their absorption bands.22,23 Because
and photochemical properties in ZnTe/CdSe type II QDs and of the higher degeneracy and closely spaced energy levels in the
pave the road for their future applications in photocatalytic VB (hole effective mass is larger than the electron in ZnTe and
reactions. CdSe),22,23 the formation and recovery of the bleach signals
The synthesis of the ZnTe seed and the ZnTe/CdSe core/ mainly reflect the dynamics of the electrons in the CB levels. In
shell QDs, and the TA measurement setup are described in the addition to the bleach peaks, there is a positive absorption
Supporting Information (SI). The SEM images of the ZnTe signal in the region red of 650 nm. A similar positive TA feature
seed and ZnTe/CdSe QDs are shown in Figure S1 (SI). Their is also observed in the ZnTe seed QDs (see Figure S5a). This
diameters are estimated to be 4.6 and 5.8 nm, respectively. The positive signal is attributed to the photoinduced absorption of
thickness of the CdSe shell is then determined to be 0.6 nm. valence holes in the ZnTe core (see SI and Figure S3 for
With the addition of the shell, the stability was found to details), consistent with the assignment in similar type II QDs
dramatically improve relative to the ZnTe seed, with a shelf life in a previous report.24 The TA kinetics probed at 500 and 600
of at least several months. The ultraviolet−visible (UV−vis) nm are shown in Figure 3b,c. The kinetics at 600 nm exhibit the
absorption and emission spectra of ZnTe/CdSe QDs and the formation of bleach signal within ∼1.5 ps, while the bleach
corresponding ZnTe seeds are shown in Figure 2. The ZnTe band at 500 nm forms within ∼0.6 ps (Figure 3c). Their
seed exhibits the first exciton absorption peak at 500 nm, and recovery kinetics are nearly identical (Figure 3b). Since the
no photoluminescence is observed.14 For ZnTe/CdSe QDs, absorption band at 600 nm is at the lowest energy level and
two absorption peaks at 500 and 600 nm are observed. near the emission peak, it is assigned to the spatially indirect
Furthermore, an emission band emerges that is centered at 652 transition from the lowest energy valence band 1s hole (1sh)
nm, indicating the formation of the expected type II level in ZnTe to the lowest energy CB 1s electron level (1se) in
material.11,20,21 The energy level diagram of these ZnTe/ CdSe, denoted as T2 in Figure 2. Our effective mass
CdSe type II QDs is plotted in the inset in Figure 2. The calculations show that the lowest energy electrons and holes
fluorescence decay of the ZnTe/CdSe QDs is shown in Figure are largely localized in the CdSe shell and ZnTe core (Figure
S2. The fitting of the decay indicates the presence of a slow S4), respectively, confirming the type II band alignment. The
2053 dx.doi.org/10.1021/jz3008886 | J. Phys. Chem. Lett. 2012, 3, 2052−2058
The Journal of Physical Chemistry Letters Letter

Figure 3. (a) TA spectra of ZnTe/CdSe type II QDs at indicated delay times after 400 nm excitation. (b) TA kinetics of the QDs probed at
indicated wavelengths (transitions) and (c) their extended view in the early time window. The kinetics of the T1 transition were obtained by
subtraction (see main text for the details). The black solid lines are the fits to eqs S1−S3.

Figure 4. (a) TA spectra of free ZnTe/CdSe QDs at indicated delay times after 585 nm excitation. (b) TA kinetics of the QDs probed at indicated
wavelengths (transitions) and (c) their extended view in the early time window.

1se and 1sh energy levels are calculated to be −3.1 and −5.1 eV, 1se in CdSe and the measured transition energy (500 nm), the
respectively, as shown in the inset in Figure 2, indicating a value energy level in the valence band for the transition at 500 nm is
of 2.0 eV for transition T2. This value agrees well with the estimated to be −5.6 eV, which is very close to the valence-
measured value of 2.07 eV at the absorption band of 600 nm, band level of the CdSe shell. We therefore assign the
supporting our assignment for T2. absorption band at 500 nm to the transition from the lowest
The TA measurements on identical QD samples were also energy delocalized valence-band hole level to the 1se level in the
carried out with an excitation at 585 nm, where only transition CdSe shell, as shown for the transition T3 in Figure 2.
T2 at 600 nm is excited. The TA spectra at different delay times Another TA feature that is significantly different at different
after 585 nm excitation are shown in Figure 4a, still showing excitation wavelengths is the ratio of the bleach amplitudes at
bleaches of the 500 and 600 nm absorption bands. The kinetics 500 and 600 nm bands (ΔA500 nm/ΔA600 nm). With 585 nm
probed at these two bands are almost identical (Figure 4b), and excitation, the ratio is ∼0.85 over the whole time window (see
the bleach signals form faster (within 0.3 ps) (Figure 4c) than Figure 4a). However, the ratio increases to ∼1.1 with 400 nm
those using 400 nm excitation. These results indicate that the excitation in the early time (within the first 2 ps), and then
transitions at 500 nm involve the same 1se level in the CdSe reduces to ∼0.85 (see Figure 3a), suggesting that the bleach
shell as T2 at 600 nm, and the electronic occupation of the level band at 500 nm has a contribution from another transition
is through excitation of T2. From the calculated energy level of (namely T1) when excited at 400 nm. The kinetics of T1
2054 dx.doi.org/10.1021/jz3008886 | J. Phys. Chem. Lett. 2012, 3, 2052−2058
The Journal of Physical Chemistry Letters Letter

Figure 5. TA spectra of ZnTe/CdSe−AQ complexes at indicated delay times after (a) 400 nm excitation and (b) 585 nm excitation. (c) TA kinetics
probed at 680 nm with 400 and 585 nm excitations and (d) their extended view in the early time window. The black solid lines are the fits according
to eqs S4 and S5.

(ΔA(t)T1) can be obtained by subtracting the T3 contribution The fit of the kinetics in Figure 3b,c according to eqs S1−S3
(ΔA(t)T3): ΔA(t)T1 = ΔA(t)500nm − ΔA(t)T3, with ΔA(t)500nm yields an internal ET time constant (1/kIET) of 0.67 ps from the
= ΔA(t)T1+T3 and ΔA(t)T3 = 0.85ΔA(t)T2 at 600 nm. The ZnTe core to the CdSe shell. This internal charge separation
kinetics of T1 are also plotted in Figure 3b,c. The T1 time is similar to reported values in CdTe/CdSe10 and ZnSe/
contribution is only present with a high energy excitation and CdS25 type II nanocrystals. The internal charge recombination
can be attributed to the transition from the 1sh level in ZnTe to dynamics are stretched over a long time scale from picoseconds
a delocalized conduction-band electron level, as shown in the to tens of nanoseconds as observed in TA kinetics (Figure 3b)
inset in Figure 2. On the basis of the calculated 1sh level in the and fluorescence lifetimes (Figure S2), with an averaged time
ZnTe core and the measured transition energy (500 nm), the constant of 115 ps. It is much slower than the recombination
energy level in the CB involved in the T1 transition is estimated kinetics in the ZnTe seed, whose time constant is 23 ps (Figure
to be at −2.6 eV. S5), consistent with the slow electron−hole recombination rate
The correlation between the kinetics at T1, T2, and T3 expected in the type II core/shell QDs.
transitions indicates the presence of an internal charge To study the photoinduced external charge separation and
separation process from the ZnTe core to the CdSe shell. recombination dynamics from ZnTe/CdSe type II QDs, we
With 400 nm excitation (see Figure 3c), the bleach of the T1 chose AQ as the electron acceptor. AQ molecules have been
band forms within ∼0.3 ps, due to the formation of excitons in utilized as electron acceptors to probe the charge separation
the ZnTe core. It quickly decays within ∼2 ps and leads to the and recombination dynamics in other QDs.10,26 The reduction
growth of the bleach signals of T2 and T3, indicating an potential of AQ is −4.0 eV (vs Vac.), indicating that the ET
internal ET process from the ZnTe core to the 1se level in the from ZnTe/CdSe to AQ is energetically allowed. The
CdSe shell. The kinetics at 500 nm have contributions from preparation of the QD−AQ complexes is described in the SI.
both T1 and T3 transitions, and therefore exhibit faster bleach The UV−vis absorption spectrum of the ZnTe/CdSe-AQ
formation kinetics than T2 at 600 nm. The contribution from complex is shown in Figure S6. The average number of the
T1 becomes negligible after 2 ps, and then the kinetics at 500 adsorbed AQ molecules per QD is estimated to be 1−2 (see
and 600 nm become identical (see Figure 3b), reflecting the Figure S6 for the details). The fluorescence of the QDs is
same electron depopulation process from the 1se level in the dramatically quenched by AQ.
CdSe shell through electron hole recombination and/or The TA spectra of this QD−AQ complex at different delay
relaxation to defect states. With excitation at 585 nm, the times after 400 and 585 nm excitations are shown in Figure 5a
transitions T1 and T3 are hardly excited, and the filling of and b, respectively. Unlike the TA spectra of free QDs, a broad
electrons at the 1se level in the CdSe shell is through exclusive induced absorption band centered at ∼680 nm was observed
excitation of T2, which generates the bleach signals at both T2 with a signal amplitude almost 2-fold larger than free QDs. This
and T3 bands (Figure 4). positive absorption is attributed to the formation of anion
2055 dx.doi.org/10.1021/jz3008886 | J. Phys. Chem. Lett. 2012, 3, 2052−2058
The Journal of Physical Chemistry Letters Letter

radical AQ¯ following the external charge separation. This external ET time (1/kEET) from the CdSe shell to AQ to be
assignment is consistent with the reported absorption spectrum <0.2 ps (within the limit of our instrument response function).
of AQ¯ in the literature,10,26 and confirms the occurrence of A similar external ET time was observed from a sample with 3-
external charge separation that produces a charge-separated fold less AQ, in agreement with the above estimation of 1−2
state ZnTe+/CdSe−AQ¯ with a hole in the ZnTe core and an AQs per QD (on average) in the presented data. The external
electron on AQ. Another TA feature in QD−AQ complexes ET kinetics (kEET > ∼5 ps−1) is at least 3 times faster than the
that is considerably different than free QDs is the depletion of internal ET kinetics (kIET = ∼1.4 ps−1), and hence the
the bleach signals at 500 and 600 nm bands (more than 5-fold population of ZnTe+/CdSe¯−AQ can be significantly sup-
less compared to free QDs). For free QDs, the TA signals at pressed in the charge separation process. This explains the
500 and 600 nm (T2 and T3 transitions) are mainly due to the observed depletion of bleach signals at the 500 (T3) and 600
filling of electrons in the 1se level in the CdSe shell after (T2) nm absorption bands in the TA spectra (Figure 5a,b).
internal ET. However, for the QD−AQ complex, the depletion Fitting the kinetics with 400 nm excitation yields a charge
of the bleach signals at T2 and T3 bands imply an ultrafast separation time of 0.71 ps, which is very close to the internal
external ET process. The remaining TA spectral signals at 500 ET time (0.67 ps), suggesting that the external charge
and 600 nm bands contain contributions mainly from the Stark- separation process is limited by the internal ET step when
effect-induced TA features.10,26−29 initiated from the ZnTe core. The averaged external charge
The Stark effect is generated in the charge separated state recombination time from AQ¯ to the hole in ZnTe/CdSe is
(ZnTe+/CdSe−AQ¯) where the electron on AQ and hole in found to be about 300 ps (1/kER) at both excitations.
the ZnTe core produce a local electric field and hence modulate The ultrafast external ET dynamics (<0.2 ps) from ZnTe/
the absorption spectrum of QDs, inducing corresponding CdSe type II QDs to AQs is at least 10-fold faster than those
bleach signals. This effect has been broadly observed and for CdTe/CdSe type II QDs (2.7 ps)10 or CdSe type I QDs
reported in bare CdSe QDs with surface trapped charges or in (3.4 ps)26 to the same AQ molecules with similar molecule-to-
an externally applied DC field,27−29 as well as CdSe/ZnS−AQ QD ratios. This can be caused by the high CB energy level of
and CdTe/CdSe−AQ ET systems.10,26 Furthermore, the ZnTe and the thin CdSe shell in ZnTe/CdSe type II QDs,
excitation of T1 with 400 nm light produces a shift of the which lead to a larger driving force (−ΔG) for the external ET.
bleach peak from 490 nm to 475 nm on the time scale of the A significantly enhanced ET rate induced by an increased
internal and external charge separation process in <2 ps (see driving force has been broadly observed in different ET
Figure 5a). This is due to the formation of the Stark effect and systems.30−32 For example, it was reported that an increase of
AQ¯ positive signal as the charge separation process proceeds. driving force from ca. −0.3 eV to ca. −0.9 eV (tuned by the size
With 585 nm excitation, the Stark-effect-induced bleach appears of QDs) led to a nearly 3 orders of magnitude enhancement in
at 475 nm at very early times due to the ultrafast external ET the ET rate from CdSe QDs to TiO2 nanoparticles.31 In the
that occurs in <0.2 ps (see Figure 5b). Similarly, a Stark effect studied ZnTe/CdSe−AQ complexes, the driving force for the
should also be present in ZnTe/CdSe QDs due to photoexcited external ET is calculated to be −0.9 eV, based upon the energy
charge carriers (a hole in ZnTe core and an electron in CdSe difference between the 1se level of the CdSe shell (−3.1 eV)
shell). However, the TA signals are dominated by the state- and the reduction potential of AQ (−4.0 eV). For comparison,
filling-induced features and the Stark-effect-induced signals are the driving force for external ET from the compared CdTe/
relatively much smaller and therefore not observed in the TA CdSe type II QDs to AQ (with an external ET time of 2.7 ps) is
spectra shown in Figures 3 and 4. only ca. −0.2 eV.10 An order of magnitude improvement in the
The TA kinetics probed at 680 nm (absorption of AQ¯) with ET dynamics (0.2 ps in ZnTe/CdSe vs 2.7 ps in CdTe/CdSe)
400 and 585 nm excitations are plotted and compared in Figure should therefore be caused by the large increase in the driving
5c,d. These kinetics indicate the external charge separation force.
(signal formation) and recombination (signal decay) processes. It should also be noted that the external charge
The whole charge separation process in the QD−AQ complex recombination dynamics (∼300 ps) from AQ¯ to the hole in
can be understood from the following reaction: the studied ZnTe/CdSe QDs is 2 orders of magnitude faster
than that (92 ns) in the compared CdTe/CdSe type II QDs.10y
This is caused by the thinner CdSe shell in ZnTe/CdSe QDs.
The electron on AQ recombines with the hole confined in the
core by tunneling through the CdSe shell. Therefore, the charge
recombination dynamics are mainly determined by the
thickness of the CdSe shell, which is 0.6 nm in the studied
ZnTe/CdSe QD and 1.6 nm in the compared CdTe/CdSe
With 400 nm excitation, the external charge separation starts QDs. While considering the practical applications of these type
from the ZnTe core, including both internal ET (from ZnTe to II materials in photodriven reductions, a thin shell should be
CdSe) and external ET (from CdSe to AQ) steps. However, preferred for the fast regeneration of the holes in the core to
with 585 nm excitation, the external charge separation process quickly bring the materials back to a neutral state. We note that
starts from the CdSe shell and includes only the external ET the CdSe shell thickness can be adjusted to permit longer
step. As shown in Figure 5d, the kinetics using 585 nm charge recombination times if the specific photocatalytic
excitation are faster than those observed using 400 nm process requires more time for the charge separated state to
excitation, indicating a faster charge separation process. These be available.
kinetics are similar in the decay region (Figure 5c) because they The above-discussed ultrafast charge separation and
follow the same charge recombination process from AQ¯ to the recombination dynamics indicate attractive advantages in the
hole in the ZnTe core. These kinetics are well fit according to ZnTe/CdSe core/shell type II QDs. First, the ZnTe/CdSe
eq S4. Fitting the kinetics with 585 nm excitation yields the QDs still maintain the benefits from the type II energy
2056 dx.doi.org/10.1021/jz3008886 | J. Phys. Chem. Lett. 2012, 3, 2052−2058
The Journal of Physical Chemistry Letters Letter

alignment. They have a long charge carrier lifetime relative to (4) Sitt, A.; Della Sala, F.; Menagen, G.; Banin, U. Multiexciton
the ZnTe type I seeds, ultrafast charge separation dynamics, Engineering in Seeded Core/Shell Nanorods: Transfer from Type-I to
and red-extended absorption spectra. More importantly, the Quasi-Type-II Regimes. Nano Lett. 2009, 9, 3470−3476.
ZnTe/CdSe QDs accomplish this while having a higher (5) Oron, D.; Kazes, M.; Banin, U. Multiexcitons in Type-II Colloidal
Semiconductor Quantum Dots. Phys. Rev. B 2007, 75, 035330/1−7.
reductive potential compared to other type II heterostructures.
(6) Smith, A. M.; Mohs, A. M.; Nie, S. Tuning the Optical and
Furthermore, the higher reductive potential of ZnTe enables a Electronic Properties of Colloidal Nanocrystals by Lattice Strain. Nat.
greater range of tunability in the core−shell structure by Nanotechnol. 2009, 4, 56−63.
changing the size of the ZnTe core and the thickness of the (7) Lee, D. C.; Robel, I.; Pietryga, J. M.; Klimov, V. I. Infrared-Active
CdSe shell. For example, our TA studies show that the Heterostructured Nanocrystals with Ultra Long Carrier Lifetimes. J.
extraordinarily high CB energy level of ZnTe core allows for a Am. Chem. Soc. 2010, 132, 9960−9962.
very thin CdSe shell (0.6 nm) in order to achieve efficient type (8) He, J.; Zhong, H. Z.; Scholes, G. D. Electron-Hole Overlap
II energy alignment. The observed 10-fold faster external ET in Dictates the Hole Spin Relaxation Rate in Nanocrystal Hetero-
the ZnTe/CdSe QD compared to other type II QDs can be structures. Phys. Rev. Lett. 2010, 105, 046601/1−4.
attributed to its thin CdSe shell (large quantum confinement), (9) Peng, P.; Milliron, D. J.; Hughes, S. M.; Johnson Justin, C.;
which leads to a stronger electronic coupling (a larger electron Alivisatos, A. P.; Saykally Richard, J. Femtosecond Spectroscopy of
Carrier Relaxation Dynamics in Type II CdSe/CdTe Tetrapod
density) and a much higher electron energy level in the CB.
Heteronanostructures. Nano Lett. 2005, 5, 1809−1813.
This leads to an overall larger driving force for ET and hence a (10) Zhu, H.; Song, N.; Lian, T. Wave Function Engineering for
faster ET rate. Ultimately, the stronger photodriven reductive Ultrafast Charge Separation and Slow Charge Recombination in Type
ability brought about by the higher CB energy level in the II Core/Shell Quantum Dots. J. Am. Chem. Soc. 2011, 133, 8762−
ZnTe/CdSe QD, combined with the fast rate constants for ET, 8771.
should enable the ZnTe/CdSe QDs to reduce a greater range (11) Dooley, C. J.; Dimitrov, S. D.; Fiebig, T. Ultrafast Electron
of materials that are not energetically reducible by other type II Transfer Dynamics in CdSe/CdTe Donor−Acceptor Nanorods. J.
heterostructures upon photoillumination. These dynamical Phys. Chem. C 2008, 112, 12074−12076.
studies strongly support the use of ZnTe nanomaterials for (12) Pearton, S. J. Wide Bandgap Semiconductors; William Andrew
photocatalysis, photovoltaics, and solar energy conversion Publishing: New York, 2000; p 9.
applications. (13) Schrier, J.; Demchenko, D. O.; Wang, L. W.; Alivisatos, A. P.


OpticalProperties of ZnO/ZnS and ZnO/ZnTe Heterostructures for
Photovoltaic Applications. Nano Lett. 2007, 7, 2377−2382.
ASSOCIATED CONTENT (14) Zhang, J.; Jin, S. Y.; Fry, H. C.; Peng, S.; Shevchenko, E.;
*
S Supporting Information Wiederrecht, G. P.; Rajh, T. Synthesis and Characterization of
Synthesis and experimental setup, SEM images, fluorescence Wurtzite ZnTe Nanorods with Controllable Aspect Ratios. J. Am.
decay, electron and hole distribution function, details of TA Chem. Soc. 2011, 133, 15324−15327.
analysis, and UV−vis absorption spectrum. This material is (15) Zhang, J.; Sun, K.; Kumbhar, A.; Fang, J. Y. Shape-Control of
available free of charge via the Internet at http://pubs.acs.org. ZnTe Nanocrystal Growth in Organic Solution. J. Phys. Chem. C 2008,


112, 5454−5458.
(16) Bang, J.; Park, J.; Lee, J. H.; Won, N.; Nam, J.; Lim, J.; Chang, B.
AUTHOR INFORMATION Y.; Lee, H. J.; Chon, B.; Shin, J.; et al. ZnTe/ZnSe (Core/Shell) Type-
Corresponding Author II Quantum Dots: Their Optical and Photovoltaic Properties. Chem.
*E-mail: rajh@anl.gov (T.R.); wiederrecht@anl.gov (G.P.W.). Mater. 2010, 22, 233−240.
(17) Kaniyankandy, S.; Rawalekar, S.; Verma, S.; Ghosh, H. N.
Author Contributions Ultrafast Hole Transfer in CdSe/ZnTe Type II Core−Shell

These authors contributed equally. Nanostructure. J. Phys. Chem. C 2011, 115, 1428−1435.
Notes (18) Wang, C.-H.; Chen, C.-W.; Wei, C.-M.; Chen, Y.-F.; Lai, C.-W.;
The authors declare no competing financial interest. Ho, M.-L.; Chou, P.-T. Resonant Energy Transfer between CdSe/ZnS


Type I and CdSe/ZnTe Type II Quantum Dots. J. Phys. Chem. C
ACKNOWLEDGMENTS 2009, 113, 15548−15552.
(19) Rawalekar, S.; Kaniyankandy, S.; Verma, S.; Ghosh, H. N. Effect
S.J. acknowledges support for a postdoctoral fellowship from of Surface States on Charge-Transfer Dynamics in Type II CdTe/
the ANSER Center, an Energy Frontier Research Center ZnTe Core-Shell Quantum Dots: A Femtosecond Transient
funded by the U.S. Department of Energy, Office of Science, Absorption Study. J. Phys. Chem. C 2011, 115, 12335−12342.
Office of Basic Energy Sciences, under Award Number DE- (20) Kumar, S.; Jones, M.; Lo, S. S.; Scholes, G. D. Nanorod
SC0001059. Use of the Center for Nanoscale Materials, and Heterostructures Showing Photoinduced Charge Separation. Small
support of J.Z., R.S., T.R., and G.P.W., is funded by the U.S. 2007, 3, 1633−1639.
Department of Energy, Office of Science, Office of Basic Energy (21) Hewa-Kasakarage, N. N.; Kirsanova, M.; Nemchinov, A.;
Schmall, N.; El-Khoury, P. Z.; Tarnovsky, A. N.; Zamkov, M. Radiative
Sciences through Contract No. DE-AC02-06CH11357.


Recombination of Spatially Extended Excitons in (ZnSe/CdS)/CdS
Heterostructured Nanorods. J. Am. Chem. Soc. 2009, 131, 1328−1334.
REFERENCES (22) Klimov, V. I. Optical Nonlinearities and Ultrafast Carrier
(1) Kim, S.; Fisher, B.; Eisler, H. J.; Bawendi, M. Type-II Quantum Dynamics in Semiconductor Nanocrystals. J. Phys. Chem. B 2000, 104,
Dots: CdTe/CdSe (Core/Shell) and CdSe/ZnTe (Core/Shell) 6112−6123.
Heterostructures. J. Am. Chem. Soc. 2003, 125, 11466−11467. (23) Klimov, V. I. Spectral and Dynamical Properties of Multi-
(2) Donega, C. D. Synthesis and Properties of Colloidal excitons in Semiconductor Nanocrystals. Annu. Rev. Phys. Chem. 2007,
Heteronanocrystals. Chem. Soc. Rev. 2011, 40, 1512−1546. 58, 635−673.
(3) Chuang, C. H.; Lo, S. S.; Scholes, G. D.; Burda, C. Charge (24) Zhu, H.; Song, N.; Rodriguez-Cordoba, W.; Lian, T. Wave
Separation and Recombination in CdTe/CdSe Core/Shell Nanocryst- Function Engineering for Efficient Extraction of up to Nineteen
als as a Function of Shell Coverage: Probing the Onset of the Quasi Electrons from One CdSe/CdS Quasi-Type II Quantum Dot. J. Am.
Type-II Regime. J. Phys. Chem. Lett. 2010, 1, 2530−2535. Chem. Soc. 2012, 134, 4250−4257.

2057 dx.doi.org/10.1021/jz3008886 | J. Phys. Chem. Lett. 2012, 3, 2052−2058


The Journal of Physical Chemistry Letters Letter

(25) Hewa-Kasakarage, N. N.; El-Khoury, P. Z.; Tarnovsky, A. N.;


Kirsanova, M.; Nemitz, I.; Nemchinov, A.; Zamkov, M. Ultrafast
Carrier Dynamics in Type II ZnSe/CdS/ZnSe Nanobarbells. ACS
Nano 2010, 4, 1837−1844.
(26) Zhu, H.; Song, N.; Lian, T. Controlling Charge Separation and
Recombination Rates in CdSe/ZnS Type I Core-Shell Quantum Dots
by Shell Thicknesses. J. Am. Chem. Soc. 2010, 132, 15038−15045.
(27) Colvin, V. L.; Alivisatos, A. P. Cdse Nanocrystals With a Dipole
Moment In the 1st Excited State. J. Chem. Phys. 1992, 97, 730−733.
(28) Norris, D. J.; Sacra, A.; Murray, C. B.; Bawendi, M. G.
Measurement Of the Size Dependent Hole Spectrum In Cdse
Quantum Dots. Phys. Rev. Lett. 1994, 72, 2612−2615.
(29) Sacra, A.; Norris, D. J.; Murray, C. B.; Bawendi, M. G. Stark
Spectroscopy of Cdse Nanocrystallites - The Significance of Transition
Linewidths. J. Chem. Phys. 1995, 103, 5236−5245.
(30) Huang, J.; Stockwell, D.; Huang, Z.; Mohler, D. L.; Lian, T.
Photoinduced Ultrafast Electron Transfer from CdSe Quantum Dots
to Re−Bipyridyl Complexes. J. Am. Chem. Soc. 2008, 130, 5632−5633.
(31) Robel, I.; Kuno, M.; Kamat, P. V. Size-Dependent Electron
Injection from Excited CdSe Quantum Dots into TiO2 Nanoparticles.
J. Am. Chem. Soc. 2007, 129, 4136−4137.
(32) Chakrapani, V.; Tvrdy, K.; Kamat, P. V. Modulation of Electron
Injection in CdSe−TiO2 System through Medium Alkalinity. J. Am.
Chem. Soc. 2010, 132, 1228−1229.

2058 dx.doi.org/10.1021/jz3008886 | J. Phys. Chem. Lett. 2012, 3, 2052−2058

You might also like