You are on page 1of 29

Article

A Visible Light-Near-Infrared Dual-Band Smart


Window with Internal Energy Storage
Sheng Cao, Shengliang Zhang,
Tianran Zhang, Qiaofeng Yao,
Jim Yang Lee
cheleejy@nus.edu.sg

HIGHLIGHTS
A dual-band electrochromic
window with internal energy
storage was demonstrated

The window supports three


modes of operation—bright, cool,
and dark

Ta-doped TiO2 nanocrystals are


the active electrochromic
component

A dual-band electrochromic energy storage (DEES) smart window was


demonstrated for the first time using Ta-doped TiO2 nanocrystals as the active
material. The demonstrative DEES unit can independently control the visible light
and near-infrared (solar heat) transmittance with good electrochromic
performance and delivers a high charge-storage capacity of 466.5 mA h m 2 at 150
mA m 2 of current density.

Cao et al., Joule 3, 1152–1162


April 17, 2019 ª 2018 Elsevier Inc.
https://doi.org/10.1016/j.joule.2018.12.010
Article
A Visible Light-Near-Infrared
Dual-Band Smart Window
with Internal Energy Storage
Sheng Cao,1,2 Shengliang Zhang,1,2 Tianran Zhang,1,2 Qiaofeng Yao,1 and Jim Yang Lee1,2,3,*

SUMMARY Context & Scale


The integration of electrochromic smart windows with energy storage is an Electrochromic energy storage
appealing concept for green building development. Herein, we report a dual- (EES) windows can be used to
band electrochromic energy storage (DEES) window capable of independent reduce a building’s energy
control of visible light (VIS) and near-infrared (NIR, solar heat) transmittance consumption—the electricity
with a high internal charge storage. The key design feature is the use of compo- generated by building-integrated
sition-optimized Ta-doped nano-TiO2 nanocrystals as a multifunctional active photovoltaics during the day is
material to deliver a high charge-storage capability concurrently with high used to darken the EES device to
bistability and long-cycle-life electrochromism in three distinct operating block visible light, solar heat, or
modes: (1) the VIS and NIR transparent ‘‘bright’’ mode; (2) the VIS transparent both based on climate conditions
and NIR opaque ‘‘cool’’ mode; and (3) the completely opaque ‘‘dark’’ mode. or personal preferences, with
The independent control of VIS (sunlight) and NIR (solar heat) modulation was simultaneous energy storage in
successfully demonstrated in a DEES prototype with a high charge-storage the colored state. The stored
capacity of 466.5 mA hr m2 at 150 mA m2 of current density. charges are then used to color or
bleach the same or another EES
device or to power other
INTRODUCTION
electronic products at night when
Electrochromic energy storage (EES) windows are recently developed transparent solar harvesting ceases. EES
thin-film devices that change color upon application of an external voltage while windows with dual-band
providing energy storage at the same time.1–5 The primary function of EES windows capability and high internal
in modern buildings is to provide a comfortable environment for the occupants by charge storage are, however, not
regulating the transmission of solar irradiation. The color change is caused by the yet demonstrated because of the
reduction or oxidation reaction of the electrochromic material, which also simulta- limitations of the electrochromic
neously charges or discharges the electrochromic material similar to a battery or materials investigated to date. We
capacitor operation.6,7 Hence, in addition to controlling light transmission, an EES discovered that doped nano TiO2
window can also be used to power electronic devices to improve a building’s energy can deliver the independent
efficiency.8,9 The potential for EES windows to further reduce a building’s energy control of visible light and near-
consumption is the driving force behind the current interest in academic and indus- infrared transmittance and a
trial research.8–12 sizeable energy storage function.
A demonstrative dual-band EES
Electrochromic materials are central to the design of EES windows. Many current EES (DEES) was fabricated to confirm
windows are WOx-based nanomaterials that have shown the best electrochromic perfor- the functionality and the rational
mance to date.10,11,13–16 An important consideration for the performance evaluation of design of next-generation smart
EES windows (vis-à-vis normal electrochromic windows) is the co-requirement for color- windows.
ation efficiency and charge-storage capacity.1 Coloration efficiency (CE) is defined as
CE = DOD/DQ where DOD is the optical density change enabled by DQ (the charge
per unit area delivered to the electrochromic material). In conventional electrochromic
applications, when two active materials can provide the same coloration ability (DOD),
the one with the lower charge density (DQ) is chosen to deliver a higher CE. For EES win-
dows, however, a high charge density is required for high charge-storage capacity, and

1152 Joule 3, 1152–1162, April 17, 2019 ª 2018 Elsevier Inc.


hence CE is no longer the single deciding factor.12,17,18 Since the charge density of most
WOx-based electrochromic materials is below 75.3 mA hr g1,3,19,20 it is prudent to
explore other electrochromic materials that can provide the same coloration ability
while providing a higher charge-storage capacity than WOx for a better implementation
of the EES windows in practice.

In addition to the capacity requirement for charge storage, the EES windows should
ideally support the independent control of sunlight and solar heat transmission
based on weather conditions and/or personal preferences. The selective modulation
of the visible light (VIS) and near-infrared (NIR) transmittance of the solar spectrum
has, however, not been demonstrated by the EES windows reported to date due pri-
marily to the electrochromic materials used for these windows.1,3,8–10,14,15,17,20
Recent research on the localized surface plasmon resonance (LSPR) of doped semi-
conductor nanocrystals (NCs) has inspired the material design for selective NIR
shielding.21,22 Several transparent doped NC films have shown the ability to signif-
icantly shift their LSPR peaks upon an applied potential to reduce their optical trans-
mittance in the NIR region.21,23–25 This NIR modulation effect is primarily capacitive
and can be used together with Faradaic reaction-based electrochromism to achieve
the independent control of VIS and NIR transmittance.25–29 Dual-band electro-
chromism combined with energy storage is expected to expand the functionality
of EES windows but is a concept that has not yet been explored and demonstrated.

Herein, we present our design of a dual-band electrochromic energy storage (DEES for
short) window capable of the independent control of solar heat and sunlight transmit-
tance, concomitantly with a high charge-storage capability. The design was demon-
strated by using Ta-doped TiO2 NCs as a model multifunctional material. Nano-TiO2
is a high-capacity insertion-type metal oxide that can be used for the lithium-ion batte-
ries.30,31 We have shown previously that colloidal Ta-doped TiO2 NCs have strong LSPR
absorption that is useful for NIR transmittance modulation.32 The working electrode
(WE) solution processed from the LSPR colloidal Ta-doped TiO2 NCs in this study deliv-
ered a high charge-storage capacity (183 mA hr g1 at the 1 C rate) concurrently with a
good dual-band electrochromic performance, which includes high dynamic range for
VIS and NIR light modulation (89.1% at 550 nm and 81.4% at 1600 nm), strong bistability
(transmittance changes in both the colored and bleached states are less than 1.5% over
an hour under the open circuit condition), and good cycle stability (optical modulation
losses of 0.2% at 550 nm, and 6.0% and 1,600 nm in 2,000 cycles). The Ta-doped nano-
TiO2 WE was also tested at the demonstrative prototype level by pairing it with an NiO-
based electrode to form a full-cell DEES device. The prototype demonstrated indepen-
dent control of VIS and NIR transmittance together with a high charge-storage capacity
of 466.5 mA hr m2 at 150 mA m2 of current density to confirm the viability of the DEES
concept.

RESULTS AND DISCUSSION


Preparation and Microstructural Characterization of DEES Electrodes
The preparation of the DEES electrode began with the synthesis of Ta-doped TiO2 1Department of Chemical and Biomolecular
colloidal NCs, followed by spin-coating on an ITO glass and thermal annealing in air Engineering, National University of Singapore,
Singapore 119260, Singapore
(see Experimental Procedures). The Ta-doped TiO2 colloidal NCs were synthesized
2Cambridge Centre for Advanced Research and
by a one-pot fluoride-assisted method that was slightly modified for operational expe- Education in Singapore, Singapore 138602,
diency.32,33 The Ta content (as Ta / [Ti + Ta + O]%) that yielded the optimal electrochro- Singapore
mic response under the experimental conditions was 2.45 at % of Ta (Figures S1 and S2). 3Lead Contact
The absorption spectrum in Figure 1A shows the dark blue color of the tetrachloroethy- *Correspondence: cheleejy@nus.edu.sg
lene solution of Ta-doped TiO2 NCs arising from their strong absorption of the entire https://doi.org/10.1016/j.joule.2018.12.010

Joule 3, 1152–1162, April 17, 2019 1153


Figure 1. Optical Characteristics, Morphology, and Size of Ta-Doped Nano-TiO2
(A) Absorption spectra of a Ta-doped TiO 2 NC dispersion in tetrachloroethylene. The inset shows a digital photo of the solution.
(B and C) (B) TEM image and (C) particle size distribution of Ta-doped TiO2 NCs.
(D and E) (D) Cross-sectional and (E) surface SEM view of a Ta-doped nano-TiO 2 film.
(F) XRD patterns of a Ta-doped nano-TiO 2 film before and after thermal annealing in air and the anatase TiO 2 reference (JCPDS card No. 21-1272). Scale
bars in (B), (D), and (E) represent 50, 1,000, and 100 nm, respectively.

NIR region (Figure 1A inset). The strong NIR absorption originated from the Ta5+ sub-
stitution of Ti4+ in the TiO2 crystal lattice, and the resulting increase in the free charge
carrier concentration brought about the LSPR effect.22,32 The Ta-doped TiO2 NCs
were pseudo-spherical in shape (Figure 1B) with well-resolved lattice fringes (Figure S3)
and a fairly uniform diameter distribution of 10.3 G 1.1 nm (Figure 1C) as measured by
transmission electron microscopy (TEM). For the fabrication of an optically transparent
Ta-doped nano-TiO2 thin-film DEES electrode, the NCs were spin-coated on an ITO
glass and subjected to a programmed heat treatment in air (up to 400 C) to remove
the surfactant on the NC surface. The loading of the Ta-doped TiO2 NCs on the ITO
glass was 0.3 mg cm2. Cross-sectional scanning electron microscopy (SEM) indicated
the Ta-doped nano-TiO2 coating as a 1.1-mm-thick compact layer on the ITO glass
(Figure 1D) with a uniformly textured top surface (Figure 1E). The calculated solid density
of the coating was 2.73 g cm3, which is lower than the true solid density of anatase
TiO2 (3.78 g cm3), an indication that the coating was a porous film. X-ray diffraction
(XRD) analysis of the coated film (Figure 1F) showed the preservation of the anatase crys-
tal structure of the starting NCs and minimum NC sintering growth during thermal an-
nealing (based on the primary particle size calculations by the Debye-Scherrer equa-
tion). The rendering of the Ta-doped TiO2 NCs into a uniform electrode film without
changing the structure and geometrical attributes of the starting NCs also retained
most of the nanoscale advantages (LSPR effect and a large electrochemically active sur-
face area) for the electrochemical processes in the DEES window.22,24,25

Electrochemical and Electrochromic Performance of DEES Electrodes


The electrochemical and electrochromic performance of the Ta-doped nano-TiO2
electrode was evaluated in two common electrolytes, namely 0.5M LiClO4 in

1154 Joule 3, 1152–1162, April 17, 2019


Figure 2. Electrochemical Performance of a Ta-Doped Nano-TiO2 Electrode Measured in a Three-
Electrode Cell
(A) Voltammograms of the Ta-doped nano-TiO2 film at 0.25 mV s1 in different electrolytes.
(B) Plot of I P versus n1/2 for the calculation of the effective diffusion coefficient D Li in the
electrochromic host.
(C) Charge and discharge curves for energy storage via Li + insertion/extraction reactions at the 1 C
rate (1 C = 200 mA g1 ).
(D) Charge capacity of Ta-doped nano-TiO 2 at different current densities.

propylene carbonate (PC) and 0.5M LiTFSI in tetraglyme (TG), in a standard three-
electrode electrochemical cell. The measured potentials were all converted to the
Li+/Li scale. Figure 2A shows the typical cyclic voltammograms (CV) at a scan rate
of 0.25 mV s1 wherein the redox couple shows the insertion and extraction reac-
tions with TiO2.30,34 Measurements of the redox peak separation (DEp) and the in-
tegrated charge under the peaks indicate that the Li+ insertion/extraction kinetics
was more facile in the LiClO4/PC electrolyte. The effective Li+ diffusion coefficient
(DLi) calculated from the plots of peak current (IP) versus the scan rate (n) (Figure 2B,
based on the voltammograms in Figure S4) show the classical square root depen-
dence of the Randles-Sevcik equation,35,36 which was 2.26 3 1011 for the LiClO4/
PC electrolyte and 1.71 3 1011 cm2 s1 for the LiTFSI/TG electrolyte. Based on
these results, the LiClO4/PC electrolyte was selected for the subsequent electro-
chemical tests and the development of the demonstrative DEES device. The
measured DLi value is also higher than in most solvothermally produced TiO2
films.35,36 This may be attributed to the preservation of the pristine nanocrystal
morphology (and its associated benefits) and the porosity of the electrode
film (which facilitates electrolyte diffusion). Figure 2C shows the typical
charge/discharge curves of the Ta-doped nano-TiO2 electrode at the 1 C rate
(1 C = 200 mA g1) between the typical cutoff potentials for electrochromism
(3.51.5 V). The charge curve corresponding to Li+ insertion can be divided into
three regions. The sharp potential decline from the open-circuit value to 1.8 V
in region I is characteristic of the Li+ adsorption on the NC surface,23,30 a

Joule 3, 1152–1162, April 17, 2019 1155


capacitive process that changes the free electron concentration on the NC surface
and subsequently modulates the NIR transmittance by the LSPR process. Region II
corresponds well with Li+ insertion into the TiO2 crystal structure. This is the main
charge-storage mechanism that forms two immiscible phases (the anatase phase
and the Li0.5TiO2 phase) and a potential plateau.23,30,34 Region III is represented
by the gradual potential decline after the potential plateau in region II as a result
of further insertion of Li+ into the Li0.5TiO2 NC near-surface layer.23,30 The specific
charge and discharge capacities of Ta-doped nano-TiO2 at ca. 183 mA hr g1 and
179 mA hr g1 are fairly typical of a TiO2-based lithium ion battery anode.10,30 The
specific charge capacity decreased expectedly with the C-rate, but a capacity of
70 mA hr g1 was still available at the 20 C rate (Figure 2D). The electrode charged
and discharged at this high rate still provided good control of the optical modula-
tion and a switching time of only 75 s (Figure S5). These indications of the success-
ful integration of optical performance with high energy storage capacity and rate
capability are very encouraging for the development of the proposed DEES
window.

Figure 3A shows the optical transmittance of the Ta-doped nano-TiO2 electrode at


different applied potentials and their corresponding digital photos. A potential
dependent electrochromic response that is useful for the independent control of
VIS and NIR transmittance was demonstrated, whereas an un-doped TiO2 thin film
could not support the independent control of VIS and NIR transmittance (Figure S6).
The high dynamic range of dual-band spectral modulation (89.1% at 550 nm and
81.4% at 1,600 nm) was able to support operations in three distinct modes. In the
‘‘bright’’ mode at 3.5 V, the electrode was colorless and transparent to both VIS
(93.4%) and NIR (94.4%). In the ‘‘cool’’ mode at 1.8 V, the electrode was light blue
in color and blocked 81.7% of the NIR while maintaining a good VIS transmittance
(70.1%). The ‘‘cool’’ mode operating window (3.51.8 V) agrees well with region I
in Figure 2C; and hence, the NIR transmittance decrease in the ‘‘cool’’ mode could
be associated with the increase in free charge carriers in Ta-doped TiO2, which gave
rise to LSPR absorption in the NIR region through capacitive charging.23,26,33 The
‘‘dark’’ mode was activated by Li+ intercalation into the TiO2 host at 1.5 V (corre-
sponding to regions II and III in Figure 2C).23,37 The electrode film was completely
dark and blocked almost all of VIS (95.7%) and 82.5% of NIR. For a more accurate
evaluation of the electrode performance for actual solar irradiation modulation,
the solar irradiance transmittance at different operating modes was calculated
(Figure 3B and Table 1). In the 1.8 V ‘‘cool’’ model, the Ta-doped nano-TiO2 elec-
trode could block 68.8% of the solar irradiance in the NIR region while maintaining
a high VIS transmittance of 71.2%. At the 1.5 V ‘‘dark’’ mode, almost all solar irradi-
ance (93.6%) was blocked.

The speed of electrochromic response between different operating modes was


measured from the transmittance spectra in real time. The switching speed from
the ‘‘bright’’ to the ‘‘cool’’ mode (3.51.8 V) as measured at 1,600 nm (Figure 3C)
yielded a coloration time (tc) of 11.4 s and a bleaching time (tb) of 3.6 s. The
‘‘bright-dark’’ mode transition (3.51.5 V) was measured at 550 nm, where the color-
ation time (tc) was 52.6 s and the bleaching time (tb) was 9.5 s (Figure 3D). The color-
ation efficiencies of 29.7 cm2 C1 at 550 nm and 121.2 cm2 C1 at 1,600 nm also
applied to a range of charge densities (Figure S7). Since stability is another key appli-
cation consideration, the bistability and cycle stability of the Ta-doped nano-TiO2
electrode were also measured. Bistability refers to the retention of the optical
memory in both the bleached and colored states under the open circuit condition.
Figure 3E shows the transmittance changes in an electrode film over an hour under

1156 Joule 3, 1152–1162, April 17, 2019


Figure 3. Electrochromic Performance of a Ta-Doped Nano-TiO2 Electrode Film
(A) Transmittance measurements and the corresponding digital photos.
(B–D) (B) Solar irradiance spectra at three different applied potentials (3.5 V [bright], 1.8 V [cool],
and 1.5 V [dark]). Real-time transmittance changes measured at (C) 1,600 nm in the 3.5 V1.8 V
window and (D) 550 nm in the 3.5 V1.5 V window.
(E) Transmittance changes monitored at 550 nm (black line) and 1,600 nm (red line) during V-off
after operation at 3.5, 1.8, and 1.5 V for 100 s, respectively.
(F) Transmittance spectra of a Ta-doped nano-TiO 2 electrode film before and after 2,000 cycles at
3.5 V, 1.8 V, and 1.5 V, respectively.

the open circuit condition (V-off) after the imposition of a bias for 100 s. The optical
transmittance at 550 and 1,600 nm changed by only 0.5% and 0.75% in the bleached
state and by 0.3% and 1.5% in the colored state. These numbers are significantly
lower than those reported for the (single-band) EES windows.9,14 Optical memory
retention is therefore good in both the bleached and colored states of Ta-doped
TiO2. The cycle stability of the Ta-doped nano-TiO2 electrode for charge storage
was evaluated by CV at the commonly used scan rate of 20 mV s1. The electrode
could retain 95.9% of its second-cycle stored charges even after 2,000 electrochem-
ical cycles (Figure S8), which is comparable to the state-of-the-art single-band EES
windows.18,38,39 Good optical modulation of the VIS and NIR regions was still avail-
able after 2,000 cycles (0.2% decrease in transmittance modulation at 550 nm and
6.0% at 1,600 nm; Figure 3F). The good dual-band optical modulation and high
charge-storage capacity, facile electrochemical response, and operational stability

Joule 3, 1152–1162, April 17, 2019 1157


Table 1. Integrated Optical Transmittance (T) and Solar Irradiance Transmittance (T’) of a Ta-
doped TiO2 Electrode Film in the VIS (400780 nm), NIR (7802200 nm), and Entire Solar (sol,
4002200 nm) Regions at Different Applied Potentials
Mode TVIS TNIR Tsol TVIS0 TNIR0 Tsol0
Bright (3.5V) 93.4% 94.4% 94.2% 93.9% 94.4% 93.8%
Cool (1.8V) 70.1% 18.3% 29.9% 71.2% 31.2% 53.9%
Dark (1.5V) 4.3% 17.5% 14.6% 4.1% 8.7% 6.4%
R R
TðlÞdl TðlÞjðlÞdl
The calculations were based on these equations: Tsol = R , Tsol0 = R , where TðlÞ is
dl jðlÞdl
the transmittance at wavelength of l, and jðlÞ is the solar irradiance at 1.5 air mass.

at the material level were then tested at the device level to confirm the viability of an
actual DEES design.

Optical and Charge-Storage Performance of the Demonstrative DEES Device


For the demonstration of an actual device capable of dual-band electrochromism and
high charge-storage capacity, a demonstrative DEES window was assembled as fol-
lows: a spin-coated Ta-doped nano-TiO2 on ITO working electrode, a spin-coated
NiO-based on ITO counter electrode40 (for more details about the NiO-based elec-
trode, see Figures S9 and S10 in Supplemental Information), and a 1 M LiClO4/PC
electrolyte. The operating voltage window of the device was set by the desired redox
features and transmittance changes (Figure S11). The DEES device assembled as
such delivered fast coloring and bleaching (Video S1). Figures 4A and 4B show the
optical transmittance of the demonstrative DEES device operating in the 13.5 V
voltage window and the corresponding digital photos. The electrochromic perfor-
mance of the device was similar to the Ta-doped nano-TiO2 electrode in the three-
electrode measurements. The narrow absorption bands in the DEES device could
be attributed to light absorption by the organic electrolyte, which has been reported
in the previous literature.41 NiO was selected as the counter electrode active material
because of its ability to store and release Li+. A NiO-based electrode coupled to a
conventional single-band electrochromic host would, however, not be able to sup-
port the selective regulation of VIS and NIR transmittance.19,40 A NiO-based elec-
trode is, however, colored during oxidation. The NiO-based counter electrode was
therefore colored during the device NIR regulation (Figures S12 and S13). As a result,
the dynamic range for selective NIR modulation was lower than in the three electrode
measurements. Other than that, the demonstrative DEES device exhibited good
charge-storage performance. Figure 4C shows that the demonstrative DEES device,
with capacities of 466.5, 237.3, and 127.8 mA hr m2 at current densities of 150, 300,
and 600 mA m2, respectively, is among the top performers for current EES windows
(Table S1). The high charge-storage capacity of the demonstrative DESS device could
easily light up a commercial red LED after charging (Figure 4D and Video S2).

The following is a foreseeable use of a DEES smart window in a modern building: The
electricity generated by building-integrated photovoltaic (PV) is used to darken the
DEES device or attenuate the passage of solar heat (NIR irradiation) based on the
user preference. The darkening of the DEES device also simultaneously charges
the DEES device, where the stored charges can be used to color/bleach the same
or another DEES device or power other electronics products (mobile phones or
smart sensors) when the solar energy harvesting stops at night.

Conclusions
In this study, we showed that Ta-doped nano-TiO2 NCs can be used as the active
material for the construction of a viable DEES window. The solution-processed

1158 Joule 3, 1152–1162, April 17, 2019


Figure 4. Optical and Charge-Storage Performance of a Demonstrative DEES Device
(A and B) (A) Optical transmittance spectra and (B) digital photos of the DEES device between 1 and
3.5 V with a 1-mm scale bar (white line). A video showing this demonstrative device in operation is
provided in Video S1.
(C) Galvanostatic charging curves of the DEES device at different current densities.
(D) An optical image of a DEES device (in the dark state) powering a red LED. A video
demonstrating this function is given in Video S2.

Ta-doped nano-TiO2 electrode preserved most of the LSPR properties of the Ta-
doped TiO2 NCs, showing good dual-band electrochromic response to support
operations in the ‘‘bright,’’ ‘‘cool,’’ and ‘‘dark’’ modes concurrently with a high
capacity for charge storage (183 mA hr g1 at the 1 C rate, 1 C = 200 mA g1).
Good bistability and cycle stability were shown for both optical performance
and charge storage. A demonstrative DEES full-cell device was also demonstrated,
which delivered a high charge-storage capacity (466.5 mA hr m2 at 150 mA m2
current density) together with independent control and modulation of the VIS (sun-
light) and NIR (solar heat) transmittance. The Ta-doped nano-TiO2-based DEES
smart window is therefore a notable upgrade of current EES windows for modern
energy-saving buildings.

EXPERIMENTAL PROCEDURES
Materials
1-octadecene (ODE) 90%, titanium ethoxide (Ti(OEt)2, technical grade), tantalum(V)
fluoride (TaF5, 98%), oleic acid (OA, 90%), oleylamine (OLA, 70%), 1-octadecanol
(ODAL 99%), lithium perchlorate (LiClO4, anhydrous, 99.99%), lithium bis(trifluorometha-
nesulfonyl) imide (Li-TFSI, 99.95 wt%), propylene carbonate (PC, anhydrous, 99.7%),
Tetramethylammonium bis(trifluoromethanesulfonyl)imide (TBA-TFSI, 97%), tetra-
chloroethylene (TCE, anhydrous, R99%), tetraglyme (TG, >99%), nickel acetate tetrahy-
drate(Ni(CH3COO)2$4H2O, 98%), titanium isopropoxide (Ti(OC3H7)4, 97%) and lithium
acetate (CH3COOLi, 99.95%) were supplied by Sigma-Aldrich and used as received.
ITO glasses (2 cm 3 3 cm, 20 U sq1) were supplied by Latech Scientific Supply Pte.

Joule 3, 1152–1162, April 17, 2019 1159


Synthesis of Colloidal Ta-Doped TiO2 NCs
The preparation of Ta-doped TiO2 NC colloidal solution was based on the fluoride-as-
sisted one-pot method.32 A minor modification was applied where TaF5 replaced the
use of a mixture of Ta salt and NH4F in the previous synthesis. In a typical synthesis,
Ti(OEt)2 (2 mmol), TaF5 (0.2 mmol), ODE (15 mL), OLA (1 mL), OA (1 mL), and ODAL
(20 mmol) were mixed in a 50-mL three-neck flask, and degassed at 120 C for
30 min. The flask was then purged with nitrogen and kept under a nitrogen atmosphere
throughout the duration of the preparation. After heating at 280 C for 60 min for the NC
growth, the reaction mixture was cooled to 60 C, and the solid product was recovered
by precipitation with acetone and re-dispersed in hexane for storage and subsequent
solution film processing. For the measurements of the NC optical properties, TCE
was used as the solvent instead since it does not absorb in the NIR region.

Fabrication of Ta-Doped Nano-TiO2 Film


A concentrated solution of Ta-doped TiO2 NCs (60 mg mL1) in hexane was spin-
coated on a clean ITO glass at 1,500 rpm for 45 s. After drying over a hot plate at
270 C for 5 min, a second layer was applied and the procedure repeated a total
of five times. The transmittance changes in the VIS region depended on the number
of spin coatings. Five spin coatings were needed to block most of the VIS transmit-
tance. The ITO glass was then annealed in air at 400 C for 45 min to decompose the
surfactant on the NCs to form a transparent coated layer.

Fabrication of NiO-Based Film


The NiO-based counter electrode films were prepared by the sol-gel spin coating
method.40 The precursor solution was a mixture of 610 mg Ni(CH3COO)2$4H2O,
134 mg CH3COOLi, and 200 mg Ti(OC3H7)4 in 50 mL ethanol. The freshly prepared pre-
cursor solution after filtration with a 0.2 mm Teflon filter was spin-coated on a clean ITO
glass at 1,500 rpm for 60 s. After drying over a hot plate at 270 C for 2 min, a second
layer was applied, and the procedure was repeated four times. Further spin coating
would cause the film to crack or become uneven. The NiO-coated ITO glass was then
annealed in air at 350 C for 60 min at a heating rate of 2 C/min.

Assembly of Dual-Band Electrochromic Energy Storage Devices


The demonstrative DEES device was assembled using the Ta-doped nano-TiO2-
coated ITO glass as the working electrode, the NiO-coated ITO glass as the counter
electrode, and 1 M LiClO4 in PC as the electrolyte. The two electrodes were assem-
bled into an optical cell facing each other and spaced by a 3 M double-sided tape
(thickness about 0.5 mm). The electrolyte was injected into the cell cavity with a
syringe, and the cell was sealed by a transparent glue.

Structure Characterizations and Optical Measurements


The material structure, morphology, and composition were characterized by X-ray
diffraction (XRD, on a D8 Advance, Bruker), high-resolution transmission electron mi-
croscopy (HRTEM, on a JEM-2100F, JEOL), and field emission scanning electron mi-
croscopy (FESEM, on a JSM 6700F, JEOL) with energy dispersive X-ray spectroscopy
(EDS, using an INCA x-act attachment, Oxford). UV-Vis-NIR spectroscopy of the NCs
was measured by an ASD Quality Spec Pro visible-NIR spectrometer and AVANTES
spectrometer (AvaSpec-ULS2048CL-EVO + AvaSpec-NIR256-1.7).

Electrochemical and Electrochromic Measurements


Half-cell measurements of electrochemical and electrochromic properties were car-
ried out in the glove box using a custom-made three-electrode spectroelectrochem-
ical cell connected to the spectrometer and a light source with fiber-optic cables. For

1160 Joule 3, 1152–1162, April 17, 2019


the standard Li+ intercalation and de-intercalation measurements, a three-electrode
configuration with a single Li foil doubled as the counter and reference electrodes
was used. 0.5 M LiClO4/PC or 0.5 M LiTFSI/TG was used as the electrolyte. The trans-
mittance of an ITO glass that had undergone a similar heat treatment as that of the
working electrode in the same electrolyte was used as the baseline. All potentials
were quoted with respect to a Li+/Li reference. Switching time (t) was defined as
the time to achieve 90% of the full modulation in the specified potential range.
Coloration efficiency (CE) was calculated by the formula: CE = DOD/DQ = log(Tb/
Tc)/DQ, where DQ is the injected charge.42

For the demonstrative DEES device measurements, the transmission spectra for
different applied voltages were measured in a glove box by the ASD Quality Spec
Pro visible-NIR spectrometer. The background transmittance in the glove box
atmosphere was used as the baseline.

SUPPLEMENTAL INFORMATION
Supplemental Information includes 13 figures, 1 table, and 2 videos and can be
found with this article online at https://doi.org/10.1016/j.joule.2018.12.010.

ACKNOWLEDGMENTS
This work was supported by the National Research Foundation (NRF), Prime
Minister’s Office, Singapore, under its Campus for Research Excellence and Techno-
logical Enterprise (CREATE) program.

AUTHOR CONTRIBUTIONS
J.Y.L. supervised this work. J.Y.L. and S.C. designed the experiments. S.C., S.Z., T.Z.,
and Q.Y. performed the experiments. J.Y.L. and S.C. co-wrote the manuscript. All
authors were involved in the results discussion and manuscript preparation.

DECLARATION OF INTERESTS
The authors declare no competing interests.

Received: October 25, 2018


Revised: November 15, 2018
Accepted: December 7, 2018
Published: January 4, 2019

REFERENCES
1. Yang, P., Sun, P., and Mai, W. (2016). 5. Xia, X., Tu, J., Zhang, Y., Wang, X., Gu, C., Zhao, all conjugated polymers for a dual function
Electrochromic energy storage devices. Mater. X.-b., and Fan, H.J. (2012). High-quality metal smart window. Energy Environ. Sci. 11, 2124–
Today 19, 394–402. oxide core/shell nanowire arrays on conductive 2133.
substrates for electrochemical energy storage.
2. Huang, Y., Zhu, M., Huang, Y., Pei, Z., Li, H., ACS Nano 6, 5531–5538. 9. Kim, Y., Shin, H., Han, M., Seo, S., Lee, W., Na,
Wang, Z., Xue, Q., and Zhi, C. (2016). J., Park, C., and Kim, E. (2017). Energy saving
Multifunctional energy storage and conversion 6. Grote, F., Yu, Z.-Y., Wang, J.-L., Yu, S.-H., and electrochromic polymer windows with a highly
devices. Adv. Mater. 28, 8344–8364. Lei, Y. (2015). Self-stacked reduced graphene transparent charge-balancing layer. Adv.
oxide nanosheets coated with cobalt–nickel Funct. Mater. 27, 1701192.
3. Cai, G., Wang, J., and Lee, P.S. (2016). Next- hydroxide by one-step electrochemical
deposition toward flexible electrochromic 10. Tong, Z., Tian, Y., Zhang, H., Li, X., Ji, J., Qu, H.,
generation multifunctional electrochromic Li, N., Zhao, J., and Li, Y. (2017). Recent
devices. Acc. Chem. Res. 49, 1469–1476. supercapacitors. Small 11, 4666–4672.
advances in multifunctional electrochromic
7. Kang, W., Lin, M.-F., Chen, J., and Lee, P.S. energy storage devices and
4. Tian, Y., Cong, S., Su, W., Chen, H., Li, Q., (2016). Highly transparent conducting photoelectrochromic devices. Sci. China
Geng, F., and Zhao, Z. (2014). Synergy of nanopaper for solid state foldable Chem. 60, 13–37.
W18O49 and polyaniline for smart electrochromic devices. Small 12, 6370–6377.
supercapacitor electrode integrated with 11. Cai, G., Eh, A.L.S., Ji, L., and Lee, P.S. (2017).
energy level indicating functionality. Nano Lett. 8. Kim, Y., Han, M., Kim, J., and Kim, E. (2018). Recent advances in electrochromic smart
14, 2150–2156. Electrochromic capacitive windows based on fenestration. Adv. Sustain. Syst. 1, 1700074.

Joule 3, 1152–1162, April 17, 2019 1161


12. Yin, X., Jennings, J.R., Tang, W., Huang, T.J., 23. Dahlman, C.J., Tan, Y., Marcus, M.A., and nanocrystals with broadly tunable plasmon
Tang, C., Gong, H., and Zheng, G.W. (2018). Milliron, D.J. (2015). Spectroelectrochemical resonance absorption. J. Mater. Chem. C 6,
Large-scale color-changing thin film energy signatures of capacitive charging and ion 4007–4014.
storage device with high optical contrast and insertion in doped anatase titania nanocrystals.
energy storage capacity. ACS Appl. ACS Appl. J. Am. Chem. Soc. 137, 9160–9166. 34. Giannuzzi, R., Manca, M., De Marco, L., Belviso,
Energy Mater. 1, 1658–1663. M.R., Cannavale, A., Sibillano, T., Giannini, C.,
24. Pattathil, P., Giannuzzi, R., and Manca, M. Cozzoli, P.D., and Gigli, G. (2014). Ultrathin
13. Yang, P., Sun, P., Chai, Z., Huang, L., Cai, X., (2016). Self-powered NIR-selective dynamic TiO2(b) nanorods with superior lithium-ion
Tan, S., Song, J., and Mai, W. (2014). Large- windows based on broad tuning of the storage performance. ACS Appl. Mater.
scale fabrication of pseudocapacitive glass localized surface plasmon resonance in Interfaces 6, 1933–1943.
windows that combine electrochromism and mesoporous ITO electrodes. Nano Energy 30,
energy storage. Angew. Chem. Int. Ed. 53, 242–251. 35. Wang, H.-Y., Chen, H.-Y., Hsu, Y.-Y., Stimming,
11935–11939. U., Chen, H.M., and Liu, B. (2016). Modulation
25. Llordés, A., Garcia, G., Gazquez, J., and of crystal surface and lattice by doping:
14. Cai, G., Darmawan, P., Cheng, X., and Lee, P.S. Milliron, D.J. (2013). Tunable near-infrared and achieving ultrafast metal-ion insertion in
(2017). Inkjet printed large area multifunctional visible-light transmittance in nanocrystal-in- anatase TiO2. ACS Appl. Mater. Interfaces 8,
smart windows. Adv. Energy Mater. 7, 1602598. glass composites. Nature 500, 323–326. 29186–29193.
15. Li, H., Li, J., Hou, C., Ho, D., Zhang, Q., Li, Y., 26. Barawi, M., De Trizio, L., Giannuzzi, R.,
and Wang, H. (2017). Solution-processed Veramonti, G., Manna, L., and Manca, M. 36. Yue, J., Suchomski, C., Voepel, P., Ellinghaus,
porous tungsten molybdenum oxide (2017). Dual band electrochromic devices R., Rohnke, M., Leichtweiss, T., Elm, M.T., and
electrodes for energy storage smart windows. based on Nb-doped TiO2 nanocrystalline Smarsly, B.M. (2017). Mesoporous niobium-
Adv. Mater. Technol. 2, 1700047. electrodes. ACS Nano 11, 3576–3584. doped titanium dioxide films from the
assembly of crystalline nanoparticles: study on
16. Azam, A., Kim, J., Park, J., Novak, T.G., Tiwari, 27. Heo, S., Kim, J., Ong, G.K., and Milliron, D.J. the relationship between the band structure,
A.P., Song, S.H., Kim, B., and Jeon, S. (2018). (2017). Template-free mesoporous conductivity and charge storage mechanism.
Two-dimensional WO3 nanosheets chemically electrochromic films on flexible substrates from J. Mater. Chem. A 5, 1978–1988.
converted from layered WS2 for high- tungsten oxide nanorods. Nano Lett. 17, 5756–
performance electrochromic devices. Nano 5761. 37. Boschloo, G., and Fitzmaurice, D. (1999).
Lett. 18, 5646–5651. Electron accumulation in nanostructured TiO2
28. Zhang, S., Cao, S., Zhang, T., Fisher, A., and (anatase) electrodes. J. Phys. Chem. B 103,
17. Tong, Z., Liu, S., Li, X., Mai, L., Zhao, J., and Li, Y. Lee, J.Y. (2018). Al3+ intercalation/de- 7860–7868.
(2018). Achieving rapid Li-ion insertion kinetics intercalation-enabled dual-band
in TiO2 mesoporous nanotube arrays for electrochromic smart windows with a high 38. Chen, Y., Wang, Y., Sun, P., Yang, P., Du, L., and
bifunctional high-rate energy storage smart optical modulation, quick response and long Mai, W. (2015). Nickel oxide nanoflake-based
windows. Nanoscale 10, 3254–3261. cycle life. Energy Environ. Sci. 11, 2884–2892. bifunctional glass electrodes with superior
cyclic stability for energy storage and
18. Mishra, S., Yogi, P., Sagdeo, P.R., and Kumar, R. 29. Zhang, S., Cao, S., Zhang, T., Yao, Q., Fisher, A., electrochromic applications. J. Mater. Chem. A
(2018). TiO2–Co3O4 core–shell nanorods: and Lee, J.Y. (2018). Monoclinic oxygen- 3, 20614–20618.
bifunctional role in better energy storage and deficient tungsten oxide nanowires for
electrochromism. ACS Appl. Energy Mater. 1, dynamic and independent control of near- 39. Li, K., Shao, Y., Liu, S., Zhang, Q., Wang, H., Li,
790–798. infrared and visible light transmittance. Mater. Y., and Kaner, R.B. (2017). Aluminum-ion-
Horiz. 5, 291–297. intercalation supercapacitors with ultrahigh
19. Li, H., McRae, L., Firby, C.J., Al-Hussein, M., and areal capacitance and highly enhanced cycling
Elezzabi, A.Y. (2018). Nanohybridization of 30. Fröschl, T., Hörmann, U., Kubiak, P., Kucerová, stability: power supply for flexible
molybdenum oxide with tungsten G., Pfanzelt, M., Weiss, C.K., Behm, R.J., electrochromic devices. Small 13, 1700380.
molybdenum oxide nanowires for solution- Hüsing, N., Kaiser, U., Landfester, K., et al.
processed fully reversible switching of energy (2012). High surface area crystalline titanium 40. Zhou, J., Luo, G., Wei, Y., Zheng, J., and Xu, C.
storing smart windows. Nano Energy 47, dioxide: potential and limits in electrochemical (2015). Enhanced electrochromic
130–139. energy storage and catalysis. Chem. Soc. Rev. performances and cycle stability of NiO-based
41, 5313–5360. thin films via Li–Ti co-doping prepared by sol–
20. Wang, W.-q., Wang, X.-l., Xia, X.-h., Yao, Z.-j.,
Zhong, Y., and Tu, J.-p. (2018). Enhanced 31. Koketsu, T., Ma, J., Morgan, B.J., Body, M., gel method. Electrochim. Acta 186, 182–191.
electrochromic and energy storage Legein, C., Dachraoui, W., Giannini, M.,
performance in mesoporous WO3 film and its Demortière, A., Salanne, M., Dardoize, F., et al. 41. Llordés, A., Wang, Y., Fernandez-Martinez, A.,
application in a bi-functional smart window. (2017). Reversible magnesium and aluminium Xiao, P., Lee, T., Poulain, A., Zandi, O., Saez
Nanoscale 10, 8162–8169. ions insertion in cation-deficient anatase TiO2. Cabezas, C.A.S., Henkelman, G., and Milliron,
Nat. Mater. 16, 1142–1148. D.J. (2016). Linear topology in amorphous
21. Garcia, G., Buonsanti, R., Llordes, A., metal oxide electrochromic networks obtained
Runnerstrom, E.L., Bergerud, A., and Milliron, 32. Cao, S., Zhang, S., Zhang, T., and Lee, J.Y. via low-temperature solution processing. Nat.
D.J. (2013). Near-infrared spectrally selective (2018). Fluoride-assisted synthesis of plasmonic Mater. 15, 1267–1273.
plasmonic electrochromic thin films. Adv. Opt. colloidal Ta-doped TiO2 nanocrystals for near-
Mater. 1, 215–220. infrared and visible-light selective 42. Barile, C.J., Slotcavage, D.J., Hou, J., Strand,
electrochromic modulation. Chem. Mater. 30, M.T., Hernandez, T.S., and McGehee, M.D.
22. Agrawal, A., Cho, S.H., Zandi, O., Ghosh, S., 4838–4846. (2017). Dynamic windows with neutral color,
Johns, R.W., and Milliron, D.J. (2018). Localized high contrast, and excellent durability using
surface plasmon resonance in semiconductor 33. Cao, S., Zhang, S., Zhang, T., Fisher, A., and reversible metal electrodeposition. Joule 1,
nanocrystals. Chem. Rev. 118, 3121–3207. Lee, J.Y. (2018). Metal-doped TiO2 colloidal 133–145.

1162 Joule 3, 1152–1162, April 17, 2019


JOUL, Volume 3

Supplemental Information

A Visible Light-Near-Infrared
Dual-Band Smart Window
with Internal Energy Storage
Sheng Cao, Shengliang Zhang, Tianran Zhang, Qiaofeng Yao, and Jim Yang Lee
Supplemental Figures and Table

Figure S1. Typical (A) XRD pattern and (B) EDS spectrum of Ta-doped TiO2 NCs. The inset shows the
corresponding chemical composition.
Figure S2. Spectral variation of the optical transmittance (ΔT) of Ta-doped nano-TiO2 films with different Ta dopant
contents at -1.5 V (vs Ag+/Ag). For the measurements, a Pt foil counter electrode, a Ag+/Ag reference electrode (in
0.1M TBA-TFSI/PC), and a 0.1 M TBA-TFSI in PC electrolyte were used.

The optimization of the Ta content in the NCs was based on the maximization of the NIR modulation. The
optical properties of Ta-doped nano-TiO2 NCs films with different Ta contents were measured in a 0.1 M electrolyte
of a large cation (tetrabutylammonium, TBA+, radius 494 pm). This large cation was unable to intercalate the TiO2
host lattice, thereby obviating electrochromism in the VIS region. 1,2 Figure S2 shows the optical transmittance (ΔT)
changes of Ta-doped nano-TiO2 films with different Ta contents in the 0 ~ -1.5 window (vs the Ag+/Ag reference).
It was found that ΔT increased as the Ta doping content increased, and maximum modulation was obtained at a
Ta content of 2.45 at %. Increase in the Ta content increased the free electron density in the NCs until electron
scattering occurred at high dopant concentrations. 3 Since high NIR modulation was desired for this study, Ta-
doped TiO2 NCs with a Ta content of 2.45 at % was selected as the composition to use for the detailed study of
dual-band electrochromism and charge storage.
Figure S3. High-resolution TEM image of a Ta-doped TiO2 NC. The inset shows its FFT patterns along the [010]
zone axis. Scale bar, 2 nm.

The Ta-doped TiO2 NC was single-crystalline and the well-resolved lattice fringes with interplanar spacings of
0.35 and 0.47 nm correspond well with the TiO2 (101) and (002) planes respectively.4,5 The fast Fourier transform
(FFT) image (inset) shows a zone axis in the [010] direction and the angle (68.3°) between the diffraction spots is
also theoretically the angle between the (101) and (002) facets of anatase TiO2.4,5 These measurements confirm
that the Ta-doped TiO2 NCs have an intact anatase TiO2 structure.
Figure S4. Voltammograms of the Ta-doped nano-TiO2 film measured in (A) 0.5 M LiTFSI/TG and (B) 0.5 M
LiCl4/PC electrolyte at different scan rates.
Figure S5. Galvanostatic charge/discharge curves (black trace) at 20 C (1C = 200 mA g -1) and the corresponding
optical response (blue trace) measured in-situ at 550 nm.
Figure S6. Transmittance of a TiO2 NC thin film electrode at three different applied potentials at 3.5, 1.8, and 1.5
V (vs Li+/Li).
Figure S7. Optical density changes (ΔOD) as a function of the charge density measured at (A) 550 nm and (B)
1600 nm. The calculated coloration efficiencies (CE) were 29.7 and 121.2 cm 2 C-1 respectively.
Figure S8. Charge storage capacity (normalized to the second cycle) as measured by cyclic voltammetry at 20
mV s-1 for 2000 cycles between 3.5 and 1.5 V (vs Li+/Li).
Figure S9. XRD pattern of the NiO-based film showing the Bunsenite cubic nickel oxide structure (JCPDS No. 4-
0835).
Figure S10. (A) Top-view and (B) cross-sectional view SEM images showing the NiO-based coating as a ∼300
nm compact layer on the ITO glass. Scale bars, 100 nm.
Figure S11. Voltammogram of the demonstrative DEES device at 10 mV s -1

Figure S11 shows the voltammogram of the demonstrative DEES device at 10 mV s -1 with similar redox
features as the three electrode measurements. Thus the operating voltage window of the device was set for 1 V
to -3.5 V.
Figure S12. Voltammogram of a NiO-based electrode measured in a three-electrode cell with 0.5 M LiClO 4/PC
electrolyte at a scan rate of 20 mV s-1.
Figure S13. Optical transmittance of the NiO-based electrode in a three-electrode cell with 0.5 M LiClO4/PC
electrolyte. (Please note that there were overlaps in the high transmittance regions of the curves at 2.5, 3, 3.2 and
3.4 V (vs. Li+/Li)).

The oxidation peak at ~3.6 V (vs. Li+/Li) in Figure S12 can be attributed to the oxidation from Ni2+ to Ni3+ upon
Li+ de-intercalation. 6-8 This oxidation process was accompanied by a visual color change from transparent to brown
and corresponding optical transmittance changes in Figure S13. When a demonstrative DEES device was
assembled with NiO as the counter electrode, the reduction of the TiO 2 electrode has to be charge balanced by
the oxidation of the NiO electrode. The oxidation of NiO also triggered a color change of the counter electrode.
This is why “cool mode” regulation was weaker in the DEES full-cell device. This finding indicates that it is
necessary to explore alternative transparent counter electrode material with a high Li + storage capacity.
Table S1. Performance comparison of EES electrodes and devices

Electrode Device

Material Optical Optical Dual- Ref.


Capacity Capacity @
modulation, modulation, band
(mA h g-1) (mA h m-2)
∆T (%) ∆T (%) function
106.6* 32*
WO3 nanoparticle film 76.2 at 633 nm 63.7 at 633 nm No 9
(at 10 mV s-1) (at 4000 mA m-2)
MoO3-W 0.71Mo0.29O3 41.9 2.33
46.7 at 500 nm 41.9 at 632.8 nm No 8
nanoparticle film (at 0.1 A g-1) (at 500 mA m-2)
WO3 film on the silver 39.9*
81.9 at 633 nm Not mentioned Not mentioned No 10
grid/PEDOT:PSS (at 1 A g-1)
66.2*
WO3/PEDOT:PSS film Not mentioned 88 at 633 nm 70 at 633 nm No 11
(at 500 mA m-2)
75.3 23.0 mA h g-1
Mesoporous WO3 film 99.5 at 633 nm 54.3 at 633 nm No 12
(at 2 A g-1) (at 1 A g-1)
Ultrathin TiO2 222
70 at 550 nm Not mentioned Not mentioned No 13
Nanorods (at 167.5 mA g-1)
TiO2 mesoporous 176 142 mA h g-1
65 at 700 nm 30.4 at 700 nm No 14
nanotube arrays (at 128 mA g-1) (at 0.25 A g-1)
Nickel-carbonate-
152.8* 257.6* 69 at 633 nm
hydroxide nanowire 85 at 500 nm No 15
(at 5 mV s-1) (at 1000 mA m-2) 45 at 700 nm
film
466.5
(at 150 mA m-2)
155.5 mA h g-1
Ta-doped TiO2 183 89.1 at 550 nm (at 0.05 A g-1) 72.7 at 550 nm This
Yes
nanocrystal film (at 200 mA g-1) 81.4 at 1600 nm 237.3 62.2 at 1600 nm work
(at 300 mA m-2)
127.8
(at 600 mA m-2)

@ Unless otherwise stated, the unit of capacitance is mA h m-2.


𝐶 ∆𝑈
* The original reported unit was Farad. For comparison, the value is converted by the equation:𝑄 = ,where Q
3.6

and C are the charge capacity (mA h) and capacitance (F), U is the voltage window.
Supplemental References

1. Barawi, M., De Trizio L., Giannuzzi R., Veramonti G., Manna L., and Manca M. (2017). Dual band
electrochromic devices based on Nb-doped TiO2 nanocrystalline electrodes. ACS Nano 11, 3576-3584.
2. Cao, S., Zhang S., Zhang T., and Lee J.Y. (2018). Fluoride-assisted synthesis of plasmonic colloidal Ta-doped
TiO2 nanocrystals for near-infrared and visible-light selective electrochromic modulation. Chem. Mater. 30,
4838-4846.
3. Agrawal, A., Cho S.H., Zandi O., Ghosh S., Johns R.W., and Milliron D.J. (2018). Localized surface plasmon
resonance in semiconductor nanocrystals. Chem. Rev. 118, 3121–3207.
4. Zhai, P., Hsieh T.Y., Yeh C.Y., Reddy K.S.K., Hu C.C., Su J.H., Wei T.C., and Feng S.P. (2015). Trifunctional
TiO2 nanoparticles with exposed {001} facets as additives in cobalt‐based porphyrin‐sensitized solar cells.
Adv. Funct. Mater. 25, 6093-6100.
5. Dai, Y., Cobley C.M., Zeng J., Sun Y., and Xia Y. (2009). Synthesis of anatase TiO2 nanocrystals with exposed
{001} facets. Nano Lett. 9, 2455-2459.
6. Wen, R.-T., Granqvist C.G., and Niklasson G.A. (2015). Anodic electrochromism for energy-efficient windows:
cation/anion-based surface processes and effects of crystal facets in nickel oxide thin films. Adv. Funct. Mater.
25, 3359-3370.
7. Zhou, J., Luo G., Wei Y., Zheng J., and Xu C. (2015). Enhanced electrochromic performances and cycle
stability of nio-based thin films via Li–Ti co-doping prepared by sol–gel method. Electrochim. Acta. 186, 182-
191.
8. Li, H., McRae L., Firby C.J., Al-Hussein M., and Elezzabi A.Y. (2018). Nanohybridization of molybdenum oxide
with tungsten molybdenum oxide nanowires for solution-processed fully reversible switching of energy storing
smart windows. Nano Energy 47, 130-139.
9. Yang, P., Sun P., Chai Z., Huang L., Cai X., Tan S., Song J., and Mai W. (2014). Large-scale fabrication of
pseudocapacitive glass windows that combine electrochromism and energy storage. Angew. Chem. Int. Ed.
53, 11935-11939.
10. Cai, G., Darmawan P., Cui M., Wang J., Chen J., Magdassi S., and Lee P.S. (2016). Highly stable transparent
conductive silver grid/pedot:pss electrodes for integrated bifunctional flexible electrochromic supercapacitors.
Adv. Energy Mater. 6, 1501882.
11. Cai, G., Darmawan P., Cheng X., and Lee P.S. (2017). Inkjet printed large area multifunctional smart windows.
Adv. Energy Mater. 7, 1602598.
12. Wang, W.-q., Wang X.-l., Xia X.-h., Yao Z.-j., Zhong Y., and Tu J.-p. (2018). Enhanced electrochromic and
energy storage performance in mesoporous WO3 film and its application in a bi-functional smart window.
Nanoscale 10, 8162-8169
13. Giannuzzi, R., Manca M., De Marco L., Belviso M.R., Cannavale A., Sibillano T., Giannini C., Cozzoli P.D.,
and Gigli G. (2014). Ultrathin TiO2(b) nanorods with superior lithium-ion storage performance. ACS Appl.
Mater. Interfaces 6, 1933-1943.
14. Tong, Z., Liu S., Li X., Mai L., Zhao J., and Li Y. (2018). Achieving rapid Li-ion insertion kinetics in TiO2
mesoporous nanotube arrays for bifunctional high-rate energy storage smart windows. Nanoscale 10, 3254-
3261.
15. Yin, X., Jennings J.R., Tang W., Huang T.J., Tang C., Gong H., and Zheng G.W. (2018). Large-scale color-
changing thin film energy storage device with high optical contrast and energy storage capacity. ACS Appl.
Energy Mater. 1 1658–1663.

You might also like