Size-controlled_quantum_dots_reveal_the_impact_of_
Size-controlled_quantum_dots_reveal_the_impact_of_
https://doi.org/10.1038/s41567-022-01639-3
Since the discovery of high-order harmonic generation (HHG) (HHG) in solids because in addition to interband transition, non-
in solids1–3, much effort has been devoted to understand its linear intraband transition is considered to be responsible for the
generation mechanism and both inter- and intraband transi- elementary excitation process behind this phenomenon. However,
tions are known to be essential1–10. However, intraband transi- the relation between nonlinear carrier generation and HHG has not
tions are affected by the electronic structure of a solid, and yet been experimentally studied, and thus, it has remained elusive
how they contribute to nonlinear carrier generation and HHG how intraband transitions correlate with the electronic structure of
remains an open question. Here we use mid-infrared laser a solid and eventually with nonlinear carrier generation and HHG.
pulses to study HHG in CdSe and CdS quantum dots, where To clarify the impact of intraband transitions on extreme nonlin-
quantum confinement can be used to control the intraband ear optical phenomena, it is necessary to study it from materials in
transitions. We find that both HHG intensity per excited vol- which the electronic band structure and thus intraband transition
ume and generated carrier density increase when the average can be freely controlled. Because the nature of the electronic bands
quantum dot size is increased from about 2 to 3 nm. We show of quantum dots (QDs) can be continuously tuned from atom-like
that the reduction in sub-bandgap energy in larger quantum discrete states to a solid-state band continuum simply by chang-
dots enhances intraband transitions, and this—in turn— ing their size, without changing the constituting elements20–23, they
increases the rate of photocarrier injection by coupling with are ideal materials to examine the role of intraband transitions in
interband transitions, resulting in enhanced HHG. carrier generation and its relation to HHG.
Transitions between the valence band (VB) and conduction Here we studied HHG in CdSe and CdS QD films. Figure 1a
band (CB) are responsible for photon absorption in a semiconduc- (dashed lines, left) illustrates a schematic of the energy levels of a
tor11. However, a strong nonlinear optical response is not necessarily QD; Eg is the QD bandgap energy and Δsub is the first sub-bandgap
governed only by such interband transitions, and thus, we need to energy. Both parameters depend on the QD diameter. Figure 1a
understand the interplay between inter- and intraband transitions. (middle) shows the absorption spectra of CdSe QDs with different
In the case that the electric field E(t) of a laser accelerates an elec- diameters (Supplementary Section I). Owing to the strong quantum
tron in the band of a solid (which constitutes an intraband transi- confinement in these small QDs, Eg is larger than in the bulk by up
tion), the temporal change in the electron wavenumber k(t) can be to several hundred millielectronvolts (the bandgap energies of bulk
described by ℏdk(t)/dt = eE(t), where e is the electron charge and ℏ CdSe and CdS are 1.75 and 2.58 eV, respectively), and the electronic
is the reduced Planck constant. Depending on the degree of change states become discrete24–27. Figure 1a (right) shows the excitation
in k, two regimes can be considered: conventional and extreme non- and detection geometry used to measure the HHG emission spec-
linear optical phenomena. In the conventional regime (that is, under tra of the CdSe QD film (Methods) and the typical transmission
resonant or near-resonant conditions), the optical field induces only electron microscopy (TEM) images. Figure 1b shows the high-order
small changes in the wavenumber, leading to a dominant contri- harmonic (HH) intensities per excited volume as a function of pho-
bution of the interband transitions12–14. In the extreme nonlinear ton energy ℏω, namely, IHHG(ℏω), for CdSe and CdS QD films under
regime where the excitation photon energy ℏω0 is much smaller excitation with linearly polarized mid-infrared (MIR) light. The
than bandgap energy Eg, efficient carrier acceleration is possible spectra extend from the visible to ultraviolet region and the peaks
without damaging the sample by excessive carrier generation. Thus, correspond to the 7th–13th orders. We find that IHHG(ℏω) tends to
it is possible to use strong long-wavelength laser fields that induce increase as the average QD diameter d increases, and it increases
large changes in the wavenumber. In this case, intraband transitions abruptly in the range from 2 to 3 nm.
play a major role in the extreme nonlinear dynamics of the system. Figure 2 provides data on CdSe (red circles) and CdS (blue
Recently, a rather unexpected role of intraband transitions was squares): it shows the integrated HH peak intensity of each order
found: carrier injection into the CB of GaAs is enhanced by cou- (Ih, where h is the harmonic order) as a function of d to clarify the
pling of inter- and intraband transitions15. Also, it has been shown QD size dependence of the HH intensity. For CdSe QDs, Ih sub-
that nonlinear-carrier-generation processes in the extreme non- stantially increases with d in the range d ≈ 1.8–3.8 nm. For example,
linear regime play an important role in the modification of opti- I7 increases by a factor of about 100 from d = 2.1 to 3.8 nm. Note
cal and electric properties of solids in the ultrafast timescales16–19. that the Eg value of CdSe QDs with d in the range of 2.1–3.8 nm
These phenomena are related to high-order harmonic generation changes from 2.6 to 2.1 eV, but the absorption spectra (Fig. 1a)
1
Institute for Chemical Research, Kyoto University, Uji, Japan. 2Center for Computational Sciences, University of Tsukuba, Tsukuba, Japan. 3Max Planck Institute
for the Structure and Dynamics of Matter, Hamburg, Germany. ✉e-mail: hirori@scl.kyoto-u.ac.jp; kanemitu@scl.kyoto-u.ac.jp
Energy
pulse
OD (normalized)
Eg 6.4 nm HH
Δsub
(QD) 1.0 3.8 nm QD films
Eg 2.8 nm
(Bulk) 2.4 nm CdSe 3.8 nm 14 nm
0.5 2.1 nm
1.8 nm
0
d 1.5 2.0 2.5 3.0 10 nm 10 nm
Photon energy (eV)
b
Harmonic order
7 9 11 13 15
CdSe d = 2.1 nm 2.8 nm 3.8 nm 14.0 nm
7th
100
9th
10–1 11th 13th
10–2 15th
IHHG (arb. units)
10–3
CdS d = 1.6 nm 2.1 nm 2.4 nm 3.5 nm
100
10–1
10–2
10–3
10–4
2 3 4 5 2 3 4 5 2 3 4 5 2 3 4 5
Photon energy (eV)
Fig. 1 | HHG in CdSe and CdS QD films. a, Energy-level diagram (left), absorption spectra of CdSe QDs with different sizes (middle) and experimental
setup including TEM images of CdSe QD films (right). The ODs are normalized by the lowest-energy exciton peaks. The TEM images of CdSe QD film
(3.8 and 14.0 nm) are shown. b, HH spectra of CdSe and CdS QD films with different average QD diameters. The plotted intensities are the measured
spectral intensity values divided by the excited volumes (that is, the number of excited QDs times the average volume of a single QD), which were
determined by dividing the absorbance at the band edge of the QD film by the absorption cross section per unit volume.
indicate that the required number of photons in the multiphoton measurements (Supplementary Sections IV and V). The QDs with
absorption process for the VB–CB transition remains almost the d ≤ 2.8 nm have very low excited carrier densities compared with
same (that is, the band edge lies in the region near h = 7 or 2.48 eV). those in larger QDs. Hence, the QD size dependence of nd has the
Although the bandgap energy for d = 2.4 nm lies closer to the 7th same tendency as that of HH intensity (Fig. 2). This implies that
resonant multiphoton absorption than that for 2.8 nm, the HH with respect to nonlinear responses such as HHG, the primary
intensities of the smaller QDs are much smaller. Also, an almost effect of a larger crystal is more efficient nonlinear carrier genera-
constant behaviour in the range d ≈ 3.8–14.0 nm (Fig. 2a, inset) is tion, and this results in larger HH intensities in the case of solids.
observed despite the difference in the order of multiphoton absorp- If laser light is applied to a semiconductor, carriers are acceler-
tion. In addition, the excitation intensity dependence shows that the ated by the electric field and their change in behaviour in real space
HHG mechanism in CdSe QDs under these excitation conditions is corresponds to an intraband transition in the Brillouin zone or
non-perturbative (Supplementary Sections II and III). These results k space. In the case of acceleration of excited carriers in a bulk crys-
show that the size dependence of Ih cannot be simply explained by tal, they pass through a continuum of states in momentum space.
the difference in the multiphoton absorption process considering However, for a sample with discrete electronic states (due to quan-
only interband transitions. tum confinement), intraband transitions are less likely to occur due
To obtain the relation between the actual carrier density gener- to energy gaps between the discrete states. Therefore, we discuss
ated by the MIR pump pulse and HHG, we measured the transient the suppression of HHG from the viewpoint of intraband transi-
absorption (TA) change in the CdSe QD films in experiments with tions: to elaborate the impact of quantum confinement on intra-
an MIR pump and a white-light probe. The schematic of the setup band transitions for HHG, we investigated the optically induced
is shown in Fig. 3a (Methods). Figure 3a also shows a typical TA electron dynamics in a simple one-dimensional dimer chain
spectrum (expressed in terms of the differential optical density model (Supplementary Section VI). Here we considered a system
ΔOD divided by optical density OD). The four graphs shown in Hamiltonian where the different sites of the chain correspond to
Fig. 3b show the TA dynamics integrated over the energy region two species of atoms, such as Cd and Se, and the coupling between
near Eg for different QD diameters. A large absorption change can the sites is represented by the nearest-neighbour hopping param-
be observed for larger QDs (d = 3.8 and 6.4 nm), whereas smaller eter. We decomposed the perturbation part of the Hamiltonian into
QDs (d = 2.4 and 2.8 nm) exhibit notably smaller changes. Figure 3c components corresponding to inter- and intraband transitions and
shows the average number of carriers per excited volume, nd, esti- calculated the HHG power spectrum I(ℏω) of a single QD.
mated from our TA data. The carrier density was determined by Figure 4a shows I(ℏω) divided by the QD volume, IHHG(ℏω), for
comparing the results of the above MIR-pump–white-light-probe different QD diameters (chain lengths). As shown in Fig. 4b, we deter-
TA measurement with those of visible-pump–white-light-probe TA mined the QD size dependence of I7 and nd by considering different
I7
components in the Hamiltonian are set to zero (Fig. 4b, blue curve,
0
and Supplementary Section VI), the intensities are substantially
10
reduced. This result indicates that the observed nonlinear responses
are determined by the size dependence of the intraband transitions.
Note that Eg increases as d becomes smaller (∝1/d2) and energy gap
CdSe
Δsub between the quantum states with quantum numbers n = 1 and
10–1
n = 2 is roughly proportional to n2/m*d2 (Extended Data Fig. 1a),
where m* is the reduced mass. To see which parameter governs the
CdS size dependence, we also evaluated the size dependence of I7 under
the assumption of a size-independent Eg. The obtained size depen-
10–2
dence (Extended Data Fig. 1b) resembles the red solid curve shown
in Fig. 4b, and this indicates that the actual increase in Eg for smaller
BG CdSe
QDs does not govern the observed size dependence. To understand
2 4 6 10 the effect of discretization of QDs on the HH intensity, we stud-
10–3 ied the influence of m* on the size dependence: Fig. 4c shows that
I9 I7 of QDs with small diameters (d ≈ 1 nm) becomes smaller as the
100 mass becomes smaller. This result is consistent with the experi-
mental results shown in Fig. 2, since the m* value of CdSe (0.1 m0)
is smaller than that of CdS (0.2 m0) (ref. 23). Moreover, the depen-
10–1 dence of I7 on Δsub, which can be changed by varying m*, is shown in
Intensity Ih (arb. units)
0.10 8
CdSe QD films 400 6
White-light probe BG
0.05 4
0
2
1.80 1.90 2.00 2.10 3 4 5 6 7
Photon energy (eV) Diameter (nm)
b
0.30
d = 2.4 nm 2.8 nm 3.8 nm 6.4 nm
0.25
0.20
ΔOD/OD
0.15 (×5)
0.10
(×30)
0.05
(×30)
0
Fig. 3 | TA measurements. a, Schematic of the experiment and typical TA spectrum of CdSe QDs obtained using an MIR pump pulse (3 μm, 0.36 TW cm–2).
b, Time evolution of ΔOD/OD near the bandgap energy for d = 2.4, 2.8, 3.8 and 6.4 nm. Here ΔOD/OD was obtained by measuring the sample
transmissivities with and without pump pulse excitation. c, Average number of carriers per excited volume, nd, estimated from the exciton amplitudes in
TA signals. The vertical and horizontal error bars represent the standard deviation calculated from the exciton amplitude as well as the diameter. The solid
curve is a guide to the eye. Since the TA signal of the sample with d = 2.4 nm was smaller than the background level, the point at d = 2.4 nm plots the upper
limit of nd estimated from the background level.
a b
Harmonic order
7 9 11 13 15 17 19
10–8
I7 (arb. units)
(intraband + interband)
10 nm
10–8 10–8
–16
10
c d
QD Bulk
100
Energy
Energy
10–4
I7 (arb. units)
10–8
0.20m0
Intraband
10–12 0.10m0
Eg Coupling
10–16
Interband Interband
10–20 m* = 0.05m0
1 10
Diameter (nm) Density of states Density of states
Fig. 4 | Calculation results. a, HH spectra per excited volume for different QD diameters (chain lengths). For this calculation, we assumed ℏω0 = 0.35 eV
and E = 11 MV cm–1, which resembles the experimental conditions. b, Diameter dependence of I7 and nd obtained by using the full model (red circles and
squares) and diameter dependence of I7 obtained by using a model excluding the intraband-transition term (blue triangles). c, Dependence of I7 on the
diameter for different reduced masses. d, Schematic describing nonlinear carrier generation via an additional path including excitation processes due to
the coupling of inter- and intraband transitions.
important implications for light-driven control of material prop- 14. Mücke, O. D., Tritschler, T., Wegener, M., Morgner, U. & Kärtner, F. X. Role
erties and also for micromachining31,32, which can be realized, for of the carrier-envelope offset phase of few-cycle pulses in nonperturbative
resonant nonlinear optics. Phys. Rev. Lett. 89, 127401 (2002).
example, by using two light sources with different wavelengths and 15. Schlaepfer, F. et al. Attosecond optical-field-enhanced carrier injection into
a tunable phase offset or different polarization states to manipulate the GaAs conduction band. Nat. Phys. 14, 560–564 (2018).
intraband transitions33. 16. Schiffrin, A. et al. Optical-field-induced current in dielectrics. Nature 493,
70–74 (2013).
17. Schultze, M. et al. Controlling dielectrics with the electric field of light.
Online content Nature 493, 75–78 (2013).
Any methods, additional references, Nature Research report- 18. Sommer, A. et al. Attosecond nonlinear polarization and light–matter energy
ing summaries, source data, extended data, supplementary infor- transfer in solids. Nature 534, 86–90 (2016).
mation, acknowledgements, peer review information; details of 19. Higuchi, T., Heide, C., Ullmann, K., Weber, H. B. & Hommelhoff, P.
author contributions and competing interests; and statements of Light-field-driven currents in graphene. Nature 550, 224–228 (2017).
20. Wang, F. et al. Exciton polarizability in semiconductor nanocrystals.
data and code availability are available at https://doi.org/10.1038/ Nat. Mater. 5, 861–864 (2006).
s41567-022-01639-3. 21. McDonald, C. R., Amin, K. S., Aalmalki, S. & Brabec, T. Enhancing high
harmonic output in solids through quantum confinement. Phys. Rev. Lett.
Received: 21 December 2021; Accepted: 13 May 2022; 119, 183902 (2017).
Published online: 23 June 2022 22. Efros, Al. L. & Efros, A. L. Interband absorption of light in a semiconductor
sphere. Sov. Phys. Semicond. 16, 772–775 (1982).
23. Brus, L. E. A simple model for the ionization potential, electron affinity, and
References aqueous redox potentials of small semiconductor crystallites. J. Chem. Phys.
1. Chin, A. H., Calderón, O. G. & Kono, J. Extreme midinfrared nonlinear
79, 5566–5571 (1983).
optics in semiconductors. Phys. Rev. Lett. 86, 3292–3295 (2001).
24. Ekimov, A. I. et al. Absorption and intensity-dependent photoluminescence
2. Ghimire, S. et al. Observation of high-order harmonic generation in a bulk
measurements on CdSe quantum dots: assignment of the first electronic
crystal. Nat. Phys. 7, 138–141 (2011).
transitions. J. Opt. Soc. Am. B 10, 100–107 (1993).
3. Schubert, O. et al. Sub-cycle control of terahertz high-harmonic
25. Murray, C. B., Norris, D. J. & Bawendi, M. G. Synthesis and characterization
generation by dynamical Bloch oscillations. Nat. Photon. 8,
of nearly monodisperse CdE (E = sulfur, selenium, tellurium) semiconductor
119–123 (2014).
nanocrystallites. J. Am. Chem. Soc. 115, 8706–8715 (1993).
4. Golde, D., Meier, T. & Koch, S. W. High harmonics generated
26. Alivisatos, A. P. Semiconductor clusters, nanocrystals, and quantum dots.
in semiconductor nanostructures by the coupled dynamics
Science 271, 933–937 (1996).
of optical inter- and intraband excitations. Phys. Rev. B 77,
27. Pietryga, J. I. et al. Spectroscopic and device aspects of nanocrystal quantum
075330 (2008).
dots. Chem. Rev. 116, 10513–10622 (2016).
5. Kuehn, W. et al. Coherent ballistic motion of electrons in a periodic potential.
28. Liu, H. et al. High-harmonic generation from an atomically thin
Phys. Rev. Lett. 104, 146602 (2010).
semiconductor. Nat. Phys. 13, 262–265 (2017).
6. Hohenleutner, M. et al. Real-time observation of interfering
29. Sivis, M. et al. Tailored semiconductors for high-harmonic optoelectronics.
crystal electrons in high-harmonic generation. Nature 523,
Science 357, 303–306 (2017).
572–575 (2015).
30. Ludwig, M. et al. Sub-femtosecond electron transport in a nanoscale gap.
7. Luu, T. T. et al. Extreme ultraviolet high-harmonic spectroscopy of solids.
Nat. Phys. 16, 341–345 (2020).
Nature 521, 498–502 (2015).
31. Lenzner, M. et al. Femtosecond optical breakdown in dielectrics. Phys. Rev.
8. Vampa, G. et al. Linking high harmonics from gases and solids. Nature 522,
Lett. 80, 4076–4079 (1998).
462–464 (2015).
32. Jürgens, P. et al. Origin of strong-field-induced low-order harmonic
9. Garg, M. et al. Multi-petahertz electronic metrology. Nature 538,
generation in amorphous quartz. Nat. Phys. 16, 1035–1039 (2020).
359–363 (2016).
33. Sanari, Y., Otobe, T., Kanemitsu, Y. & Hirori, H. Modifying angular and
10. Aversa, C. & Sipe, J. E. Nonlinear optical susceptibilities of semiconductors:
polarization selection rules of high-order harmonics by controlling electron
results with a length-gauge analysis. Phys. Rev. B 52, 14636–14645 (1995).
trajectories in k-space. Nat. Commun. 11, 3069 (2020).
11. Haug, H. & Koch, S. W. Quantum Theory of the Optical and Electronic
Properties of Semiconductors (World Scientific, 2004).
12. Boyd, R. W. Nonlinear Optics 3rd edn (Academic Press, 2008). Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in
13. Cundiff, S. T. et al. Rabi flopping in semiconductors. Phys. Rev. Lett. 73, published maps and institutional affiliations.
1178–1181 (1994). © The Author(s), under exclusive licence to Springer Nature Limited 2022
Extended Data Fig. 1 | Additional calculation results. a, Bandgap energy Eg (blue squares) and subband gap, Δsub (red circles), as a function of the QD
diameter (chain length). b, Diameter dependence of I7 obtained by assuming a size-independent Eg (green squares) and that obtained by the full model
with a size-dependent Eg (red circles). c, Dependence of I7 on Δsub for different reduced masses.
Extended Data Fig. 2 | Schematics of multiple excitation paths. In addition to the contribution of the pure interband transition terms (left), the efficient
intraband transition in larger QDs (or bulk) opens multiple excitation paths due to the nonlinear coupling between the intra- and interband transitions
(right). These additional excitation channels due to the coupling promote nonlinear carrier injection and enhance HHG in larger QDs.
Extended Data Fig. 3 | Yield ratio. Diameter dependence of the yield ratio of the 7th order for CdSe, I7/nd. The data is normalized to the value at
d = 6.4 nm. Vertical and horizontal error bars represent the standard deviation of yield ratio and that of diameter. The solid curve is a guide to the eye.
1. use such content for the purpose of providing other users with access on a regular or large scale basis or as a means to circumvent access
control;
2. use such content where to do so would be considered a criminal or statutory offence in any jurisdiction, or gives rise to civil liability, or is
otherwise unlawful;
3. falsely or misleadingly imply or suggest endorsement, approval , sponsorship, or association unless explicitly agreed to by Springer Nature in
writing;
4. use bots or other automated methods to access the content or redirect messages
5. override any security feature or exclusionary protocol; or
6. share the content in order to create substitute for Springer Nature products or services or a systematic database of Springer Nature journal
content.
In line with the restriction against commercial use, Springer Nature does not permit the creation of a product or service that creates revenue,
royalties, rent or income from our content or its inclusion as part of a paid for service or for other commercial gain. Springer Nature journal
content cannot be used for inter-library loans and librarians may not upload Springer Nature journal content on a large scale into their, or any
other, institutional repository.
These terms of use are reviewed regularly and may be amended at any time. Springer Nature is not obligated to publish any information or
content on this website and may remove it or features or functionality at our sole discretion, at any time with or without notice. Springer Nature
may revoke this licence to you at any time and remove access to any copies of the Springer Nature journal content which have been saved.
To the fullest extent permitted by law, Springer Nature makes no warranties, representations or guarantees to Users, either express or implied
with respect to the Springer nature journal content and all parties disclaim and waive any implied warranties or warranties imposed by law,
including merchantability or fitness for any particular purpose.
Please note that these rights do not automatically extend to content, data or other material published by Springer Nature that may be licensed
from third parties.
If you would like to use or distribute our Springer Nature journal content to a wider audience or on a regular basis or in any other manner not
expressly permitted by these Terms, please contact Springer Nature at
onlineservice@springernature.com