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

Chemical Engineering Journal 415 (2021) 128935

Contents lists available at ScienceDirect

Chemical Engineering Journal


journal homepage: www.elsevier.com/locate/cej

Organic thermoelectric devices with PEDOT:PSS/ZnO hybrid composites


Woongki Lee a, Sooyong Lee a, Hwajeong Kim a, b, Youngkyoo Kim a, *
a
Organic Nanoelectronics Laboratory, Department of Chemical Engineering, School of Applied Chemical Engineering, Kyungpook National University, Daegu 41566,
Republic of Korea
b
Priority Research Center, Research Institute of Environmental Science & Technology, Kyungpook National University, Daegu 41566, Republic of Korea

A R T I C L E I N F O A B S T R A C T

Keywords: Organic thermoelectric devices (OTEDs) are considered one of the most promising thermoelectric devices for
PEDOT:PSS harvesting electricity from low-temperature heat sources. Here it is demonstrated that the hybrid composites of
ZnO poly(3,4-ethylenedioxythiophene):poly(styrene sulfonate) (PEDOT:PSS) and zinc oxide (ZnO) particles enhance
Hybrid composites
the performances of OTEDs due to the improved electrical conductivity by the bridging role of ZnO particles. The
Organic thermoelectric
Flexible thermal sensors
detailed study showed that the addition of ZnO particles to the aqueous solutions of PEDOT:PSS led to the
increased viscosity owing to the specific interactions between sulfonic acid groups in PSS and hydroxyl groups
(and/or zinc cations) in ZnO. The OTEDs with the PEDOT:PSS/ZnO hybrid composites with 70 wt% of 45 μm-
sized ZnO particles exhibited 0.55 nW, which is ca. 7.8 times higher power than the devices with the pristine
PEDOT:PSS sheets (0.07 nW). The wearable thermal sensors with the PEDOT:PSS/ZnO hybrid composite films
showed fast thermoelectric responses toward heat and cool sources.

1. Introduction nanotube (CNT) [17–20].


For practical applications such as patch-type devices on hot and/or
Organic thermoelectric devices (OTEDs) have attracted significant cool sources, OTEDs should be fabricated with a vertically-stacked
interest due to their capability of converting heat to electrical energy in structure [2,3,21–25]. However, most of the previous studies with
a form of flexible shapes applicable to various heat sources including high TE performances were conducted by employing a horizontal
human body [1–5]. The advantageous characteristics of OTEDs, structure that can be simply prepared and is able to easily keep the
compared to inorganic thermoelectric devices, can be practically real­ temperature difference in the lateral direction owing to the large gap
ized by employing polymeric materials as an active layer for the charge between electrodes (see Table S1). To fabricate the vertically-stacked
generation upon temperature difference [6–8]. Here it is important for OTEDs with sufficiently thick sheets for reasonable temperature differ­
OTEDs that the polymeric active layers possess high electrical conduc­ ences, the solutions of materials should be viscous enough to deliver
tivity and low thermal conductivity for efficient charge transport and such thick organic sheets (films). Although the pristine PEDOT:PSS so­
effective thermal gradient, respectively [9–11]. lutions can be concentrated by evaporating water for a higher viscosity,
In this regard, conducting polymers such as poly(3,4- the resulting pristine PEDOT:PSS sheets show a shrink shape with poor
ethylenedioxythiophene):poly(styrene sulfonate) (PEDOT:PSS) and sheet edges and large internal cracks made by shrinkage stress and se­
poly(aniline):camphorsulfonic acid (PANI:CSA) have been introduced as lective drying between surficial and internal polymer matrix [26,27].
an active layer for OTEDs [12–20]. In particular, the PEDOT:PSS poly­ Hence, a novel idea is necessary to secure both viscosity and well-
mer was extensively applied for OTEDs because of its high electrical maintained sheet (thick film) shape for vertically-stacked OTEDs.
conductivity up to ca. 2000 S∙cm− 1 and environmentally friend process Because the PEDOT:PSS polymers exist in a form of nanoparticles in
using aqueous solutions [14–16]. It has been reported that the OTEDs aqueous solutions, microscale or nanoscale voids can be inevitably made
with the pristine PEDOT:PSS films in a horizontal (planar) structure can in the PEDOT:PSS sheets during wet-coating processes [28–30].
generate an electricity with a power factor of 334 μW m− 1K− 2 by Recently, aerogel and/or aery materials containing voids have been
applying acid/base post-treatments, 284 μW m− 1K− 2 by compositing reported to help better flexibility in organic thermoelectric devices
with tellurium, and 526 μW m− 1K− 2 by compositing with carbon because of their spatially lenient structures compared to corresponding

* Corresponding author.
E-mail address: ykimm@knu.ac.kr (Y. Kim).

https://doi.org/10.1016/j.cej.2021.128935
Received 7 October 2020; Received in revised form 6 January 2021; Accepted 7 February 2021
Available online 13 February 2021
1385-8947/© 2021 Elsevier B.V. All rights reserved.
W. Lee et al. Chemical Engineering Journal 415 (2021) 128935

densely filled ones [1–5,10,31,32]. Here it is worthy to note that these freeze-drying processes applied).
voids or aery parts are advantageous in keeping the thermal gradient
across the devices when it comes to the thermally insulative nature of
voids (filled airs or gases depending on the fabrication environment). 2.2. Measurements and characterizations
However, the voids can be critical for charge transport owing to their
electrically insulative characteristics [33,34]. Therefore, it is required to The ionization potential of composite samples was measured using a
introduce a microscale or nanoscale bridge into the voids for the photoelectron yield spectrometer (AC-2, Hitachi High Tech). A Raman
improved charge transport in the presence of maintaining the thermal spectrometer (InVia Reflex, Renishaw) was used for the measurement of
insulating performances and better flexibility. the Raman spectra for the composite samples (excitation wavelength =
In this work, we attempted to introduce zinc oxide (ZnO) particles 532 nm). A high performance X-ray photoelectron spectrometer (XPS,
into the PEDOT:PSS solutions because the extra sulfonic acid groups in ESCALAB 250, Thermo Scientific, Inc.) was used for the measurement of
the PSS units were expected to interact with the hydroxyl groups (and/ XPS spectra for the composite films. The viscosity of solutions was
or zinc cations) of ZnO particles (due to the basic nature of metallic measured using a rotational rheometer (DHR1, TA Instrument Ltd.) at
oxides). Note that ZnO was chosen among various metallic oxides 25 ◦ C at a shear rate range between 1.0 and 700 s− 1. The thermal con­
because the (crystalline) ZnO materials have in general good electron ductivity of sheets (thick films) in out-of-plane (thickness) direction was
mobility leading to a positive effect on the resulting sheets (films) measured with a laser flash method using a light flash apparatus (LFA
compared to other metallic oxides with very high electrical resistivity 467 HyperFlash, Netzsch) (standard deviation = ca. 0.001 W m− 1K− 1).
such as alumina (Al2O3) [35,36]. For the systematic investigation, three The optical microscope (OM) images were measured using an optical
different sizes (0.1, 5 and 45 μm) of ZnO particles were employed by microscope (SV-55, Sometech), while a field-emission scanning electron
varying their contents. The PEDOT:PSS solutions became viscous after microscope (FE-SEM, JSM-6701F, JEOL) was used for the measurement
adding ZnO particles, providing an effective state for making thick of the high-resolution images on the surface and cross-sectional parts of
composite sheets. The device results showed that the thermoelectric composite sheets. The thermoelectric performances were measured
performances of OTEDs were greatly enhanced for the PEDOT:PSS/ZnO using the home-built system equipped with an electrometer (Keithley
hybrid composites, compared to the pristine PEDOT:PSS sheets, due to 6517, Keithley Instruments Inc.), thermocouples, and a temperature
the improved electrical conductivity. The detailed morphology and control module (NI-9211, National Instruments, LLC). Thermal gradi­
spectroscopic investigations on the PEDOT:PSS/ZnO hybrid composites, ents were made by employing two peltier devices (10 × 10 mm2, TEC1-
which could act as a flexible thermal sensor with fast thermal responses, 0706, PTHAUS), contacting directly the backside of the Cu electrodes,
were performed with high-resolution scanning electron microscopy which were controlled by a power supply (SPD1168X, Siglent). Note
(SEM), Raman spectroscopy, and X-ray photoelectron spectroscopy that all devices before measurement were subjected to a pretreatment
(XPS). (aging process) by exposing to the condition of thermal difference (ΔT ≥
50 K) for more than 1 h. For the measurement of wearable sensor per­
2. Experimental section formances, the thermoelectric devices with the Cu foil electrodes were
worn on a human fingertip and approached toward a heat or cool source.
2.1. Composite sheet preparation and device fabrication The real time current and voltage signals from the thermoelectric de­
vices on the fingertip were measured using the same electrometer
The ZnO particles (spherical shape) with three sizes (0.1, 5, 45 μm, (Keithley 6517).
Merck) were separately added to the pristine PEDOT:PSS solutions
(Clevios PH500, Heraeus, solid content = 1.0–1.3 wt% in aqueous so­ 3. Results and discussion
lution) by varying their contents (30, 50, 70 wt%) with respect to the
solid content of PEDOT:PSS. The solutions of ZnO and PEDOT:PSS in 3.1. Device structure and hybrid composites
vials were subjected to mixing using a vortex mixer for more than 30
min, followed by stirring using a magnetic stirring bar on a hot plate As shown in Fig. 1a, the present OTEDs were fabricated by vertically
stirrer for 1 day at room temperature. Then the solutions were poured stacking the Cu electrodes in the presence of PEDOT:PSS/ZnO composite
into a Teflon mold with an inner-diameter of 61 mm and dried for 1 h at sheets between the electrodes. The PEDOT:PSS/ZnO composite sheets
60 ◦ C. This process led to a gel-like state for all the mixture solutions were prepared by casting corresponding solutions on the Teflon mold
except the pristine PEDOT:PSS solution. These gel-like mixtures were via freeze drying processes (see details in the experimental section).
subjected to freeze-drying processes at − 45 ◦ C using a freeze dryer (LP Upon the initial heat treatment at 60 ◦ C for 1 h, the mixture solutions of
20, IlshinBioBase) at a reduced pressure of 100 mTorr. The temperature PEDOT:PSS and ZnO particles became a gel state, whereas such a gel
was slowly increased from − 45 ◦ C to 0 ◦ C for 150 min and then to room state was not made for the pristine PEDOT:PSS solution (see Fig. S1).
temperature for 3 days at the same pressure. Note that the pristine The reason can be attributable to the formation of strong interactions
PEDOT:PSS sheets were prepared with the same process even though between hydroxyl groups (and/or Zn2+) in ZnO particles and sulfonic
they were not gelated but just became viscous during the heating pro­ acid groups (and/or SO−3 ) in PSS chains, which might make bridges
cess. After the freeze drying processes were finished, free-standing among the PEDOT:PSS domains leading to partial network structures
composite sheets were made (thickness >1.0 mm) and tailored into responsible for high viscosity [37,38]. As the content of ZnO particles
small pieces with a size of ca. 8 × 8 mm2. The PEDOT:PSS/ZnO com­ increased, the viscosity of mixture solutions was basically increased (see
posite sheets were sandwiched between copper (Cu) electrodes equip­ Fig. S2). The solution viscosity reached ca. 2530 mPa.s (shear rate = 10
ped in a home-built holder system, which led to a vertically-stacked s− 1) at 70 wt%, even though it was only ca. 10 mPa.s at 0 wt% (the
geometry of OTEDs. The bending test was conducted by employing a pristine PEDOT:PSS solution) (see the information in the caption of
home-built system that can change the radius of curvature such as 5, 10, Fig. S2). In detail, the gelation was partly made at 30 wt% but it was well
20 mm and infinity (flat state). The gap between the Cu electrodes was made at the higher contents (50 wt% and 70 wt%). However, more
kept 500 μm. For wearable applications, thin copper foils (thickness = addition of ZnO particles (over 80 wt%) resulted in extremely viscous
50 μm) were attached on the both sides of the PEDOT:PSS/ZnO com­ states with almost no liquid flows so that the mixture solutions with such
posite sheets (films), followed by encapsulating with a 50 μm-thick higher contents could not be used for the preparation of sheets (see
cellophane tape. For spectroscopic characterizations, the pristine and Fig. S3). Note that the pristine PEDOT:PSS sheets were obtained without
composite samples (thickness = ca. 10 μm) were prepared by dispensing any gelation step via typical solidification process in the presence of
ca. 100 μl solutions on glass substrates and dried for 1 h at 60 ◦ C (no large scale shrinkages.

2
W. Lee et al. Chemical Engineering Journal 415 (2021) 128935

Fig. 1. (a) Illustration of organic thermoelectric devices (OTEDs) with the PEDOT:PSS/ZnO hybrid sheets. (b) Photographs for the PEDOT:PSS/ZnO solutions casted
on the Teflon mold (left), leading to free-standing composite sheets (middle) and tailored composite sheets (size = 8 × 8 mm2, right). (c) A proposed scheme for the
gelation in the mixture solutions of PEDOT:PSS and ZnO particles by the interaction between ZnO and PSS units: Note that the anionic parts (SO−3 ) of PSS upon
dissociation of protons in aqueous solutions might bind to the positively charged Zn atoms.

3.2. Thermoelectric performances shown in Fig. 2a (top panel), the output current (IOUT) from the OTEDs
was almost gradually increased with the temperature gap (△T) between
Based on the above result, the ZnO content was fixed to 70 wt% for the electrodes for all the devices. Interestingly, the output current was
the investigation of ZnO size in the OTEDs with the PEDOT:PSS/ZnO higher for the OTEDs with the hybrid composites than those with the
hybrid composite sheets. The mixture solutions with three different sizes pristine PEDOT:PSS sheets. In addition, at a fixed temperature gap, the
of ZnO particles (0.1, 5 and 45 μm) exhibited almost similar flow higher output current was measured for the OTEDs with the larger size
characteristics after sequential mixing steps (vortex for 30 min and of ZnO particles (see Fig. 2b). As observed from Fig. 2a (bottom panel),
stirring for 1 day) (see Fig. S4). The OTEDs with the hybrid composites the electrical conductivity (σ) was still higher for the OTEDs with the
having the different size of ZnO particles were fabricated by stacking the hybrid composites than those with the pristine PEDOT:PSS sheets over
Cu electrodes on both sides of the tailored PEDOT:PSS/ZnO sheets. As the entire temperature range. This result may support that the ZnO

3
W. Lee et al. Chemical Engineering Journal 415 (2021) 128935

Fig. 2. (a) Current output (IOUT) and electrical conductivity (σ) as a function of temperature difference (△T) from 3 to 50 K according to the size of ZnO particles. (b)
IOUT and σ as a function of the size of ZnO particles at the representative temperature difference (△T = 10, 30, 50 K).

particles are well positioned without huge change in the PEDOT:PSS/ bottom panel in Fig. 3a). Note that a marginal deviation in the output
ZnO hybrid composites under the present temperature variation. voltage and Seebeck coefficient can be observed from Fig. 3b.
However, the trend of output voltage (VOUT) was a bit different from As a consequence, the higher output power (POUT) was obtained for
that of output current. As shown in Fig. 3a (top panel), the output the OTEDs with the larger size of ZnO particles over the entire tem­
voltage of the OTEDs with the hybrid composites was not so pronounced perature range (see top panel in Fig. 4a). The output power reached ca.
with respect to the size of ZnO particles even though it was higher than 0.55 nW for the OTEDs with the hybrid composites containing the 45 μm
that of the OTEDs with the pristine PEDOT:PSS sheets. This result im­ ZnO particles, whereas the pristine PEDOT:PSS sheets resulted in ca.
plies that the conjugation length of thiophene in PEDOT:PSS might be 0.07 nW (see Fig. 4b and Table 1). The resulting power factor for the
affected by the content of ZnO particles, not the particle size (see the OTEDs with the hybrid composites containing the 45 μm size ZnO par­
results of Raman spectra in the next section). Here it is considered that ticles reached ca. 1.7 × 10− 6 μW m− 1K− 2 at △T = 50 K (note that the
the relatively high Seebeck coefficient of ZnO materials might partly thermal conductivity of the composite sheets at 70 wt% was measured as
influence on the resulting value for the hybrid composite sheets even 0.053 W m− 1K− 1 in the thickness direction). Although the resulting
though the ZnO particles in the hybrid composite sheets did not make power factor is lower than those in the literatures for vertical devices
their own large phases as an individual component for enough voltage (see Table S1), the present OTEDs with the hybrid composite sheets have
difference [39,40]. As a result, the Seebeck coefficient (S) was almost strong potential in terms of easy control of sheet (film) thickness for
similar for the OTEDs with the hybrid composites but it was always wearable and flexible applications even though an actual limitation is
higher than that of the devices with the pristine PEDOT:PSS sheets (see expected for textile and/or sponge type materials reported in the

Fig. 3. (a) Voltage output (VOUT) and Seebeck coefficient (S) as a function of temperature difference (△T) from 3 to 50 K according to the size of ZnO particles. (b)
VOUT and S as a function of the size of ZnO particles at the representative temperature difference (△T = 10, 30, 50 K).

4
W. Lee et al. Chemical Engineering Journal 415 (2021) 128935

Fig. 4. (a) Power output (POUT) and power factor (PF) as a function of temperature difference (△T) from 3 to 50 K according to the size of ZnO particles. (b) POUT
and PF as a function of the size of ZnO particles at the representative temperature difference (△T = 10, 30, 50 K).

3.3. Morphology, work function, Raman and XPS spectra


Table 1
Summary of thermoelectrical characteristics for the OTEDs with the PEDOT:
To understand the performance change according to the size of ZnO
PSS/ZnO hybrid composites (thickness = 500 μm) according to the size of ZnO
particles (ZnO content = 70 wt%). Note that the temperature difference was 50 particles, the surface and cross-section parts of the hybrid composites
K. were examined by employing a high-resolution optical microscope. As
shown in Fig. 5a, it is obvious that the surface of hybrid composite sheets
Parameters (ΔT = 50 DZnO (μm)
K)
was quite differently formed compared to that of the pristine PEDOT:PSS
w/o ZnO 0.1 5 45 sheets. Note that some ZnO particles exist on the surfaces of sheets
|IOUT| (nA) 110.1 184.9 282.5 643.1 irrespective of their sizes (see Fig. S6 for the enlarged images). The cross-
|VOUT| (μV) 616 905 875 854 sectional images in Fig. 5b deliver very interesting stacking structures of
POUT (nW) 0.0678 0.1673 0.2472 0.5492
thin polymeric flakes, which look relatively close each other and regular
σ (μS∙cm− 1) 13.97 15.96 25.22 58.83
S (μV∙K− 1) 12.32 18.10 17.50 17.08 for the PEDOT:PSS/ZnO hybrid composites compared to the pristine
PF (μW m− 1K− 2) 2.12 × 5.23 × 7.72 × 1.72 × PEDOT:PSS sheets (note that the gaps in the stacking structures were
10− 7 10− 7 10− 7 10− 6 slightly varied depending on the positions and breaking angles). The
better stacking structures (narrower gaps) in the hybrid composites can
be attributed to the anchoring role of ZnO particles which makes the
literatures. All parameters for the present OTEDs are summarized in
PEDOT:PSS domains closer each other by the interaction between hy­
Table 1. Note that the best output power was measured when the con­
droxyl groups (and/or Zn2+) in ZnO particles and sulfonic acid groups
tent of ZnO particles was 70 wt% even though the output voltage was
(and/or SO−3 ) in PSS chains.
higher at 30 wt% and 50 wt% than 70 wt% (see Fig. S5).

Fig. 5. Optical microscope (OM) images for the PEDOT:PSS/ZnO hybrid composite sheets with three different sizes of ZnO particles (DZnO = 0.1, 5, 45 μm): (a)
Surfaces, (b) cross-sections. The dashed circles in pink indicate the ZnO particles in the hybrid composite sheets. (For interpretation of the references to colour in this
figure legend, the reader is referred to the web version of this article.)

5
W. Lee et al. Chemical Engineering Journal 415 (2021) 128935

A close look into the cross-section parts was attempted by employing that the shape of flakes were slightly deformed or damaged upon sam­
a field-emission scanning electron microscope (FE-SEM). As shown in pling for the observation of the inner region in the cross-section parts).
Fig. 6, the pristine PEDOT:PSS flakes have very smooth surfaces without As the size of ZnO particles increased, some larger aggregates were
any particles. In contrast, small particles were adhered to the surfaces of formed on the flake surfaces and between the flakes. In addition, it is
the hybrid composite flakes irrespective of the size of ZnO particles (note considered that some ZnO particles were included inside the flakes in the
hybrid composites. This result evidences that the ZnO particles were not
separately present but attached to the PEDOT:PSS domains (flakes) by
the interaction between ZnO and PSS.
Based on the morphology analysis results, it is understood that the
addition of ZnO particles led to the formation of actual hybrid mixtures
(composites) of PEDOT:PSS and ZnO in bulk and some ZnO particles
placed on the surfaces of the composite flakes. Therefore, as illustrated
in Fig. 7, the enhanced conductivity of the hybrid composites (see
Fig. 2a) can be attributed to the close stacking of flakes and the bridging
role of ZnO particles which might help charge transport across the
voided parts (gaps) between flakes. The higher electrical conductivity
with the larger ZnO particles can be explained by the better role of the
ZnO bridges (see the illustration for DZnO = 45 μm in Fig. 7).
To further understand the enhanced conductivity by the addition of
ZnO particles, the ionization potential of film samples was measured
using a photoelectron yield (PEY) spectrometer. As shown in Fig. 8a, the
ionization potential became gradually lowered with the content of ZnO
particles (45 μm). The ionization potential was ca. − 4.9 eV and − 4.73
eV for the pristine PEDOT:PSS and hybrid composite (70 wt%) films,
respectively. Here it is found that the ionization potential of ca. − 4.73
eV for the hybrid composite films (70 wt%) does well match the work
function of Cu (ca. − 4.7 eV), which might contribute to the reduction of
electrical resistances at the interfaces between the Cu electrodes and the
hybrid composites. In addition, the Raman spectra in Fig. 8b disclosed
that the symmetrical Cα-Cβ stretching peak of thiophene rings in the
PEDOT chains was shifted toward a lower wavenumber in the case of the
hybrid composite films. This result informs that the conjugation length
of thiophene rings in the PEDOT:PSS domains was increased by the
presence of ZnO particles when it comes to the more stable (lower en­
ergy) state at a lower wavenumber for the stretched chains than the
bended chains leading to a broken conjugation (see Fig. S7). Hence, the
extended conjugation length in the PEDOT chains might play an
important role in enhancing the electrical conductivity of the PEDOT:
PSS/ZnO hybrid composites, which eventually contributed to the
improved thermoelectric performances. Here, the interaction between
ZnO and PSS can be supported by the XPS spectra in Fig. 8c. First, the
Zn2p peaks were clearly measured for the PEDOT:PSS/ZnO films even
though those peaks were absolutely not found from the pristine PEDOT:
PSS films (see the left panel in Fig. 8c). In addition, the two Zn2p peaks
were measured at a higher energy part for the larger ZnO particles (5 and
45 μm) than the smaller ZnO particle (0.1 μm), which may basically
reflect the change of Zn atom environments according to the size of ZnO
particles in the hybrid composite films (see Fig. S8) [41,42]. As observed
in the right panel of Fig. 8c, the S2p peak of sulfonic acid (PSS) showed a
shift toward a lower binding energy for the PEDOT:PSS/ZnO films
compared to the pristine PEDOT:PSS film. This peak shift indicates that
some of the sulfonic acid groups in PSS were changed to sulfonate
groups by the interaction with the hydroxyl groups in ZnO particles
[43,44]. Note that S2p peaks of thiophene rings were similarly shifted by
the presence of ZnO particles, indicative of huge change in the envi­
ronment of sulfur atoms in the PEDOT chains.

3.4. Wearable thermal sensors

The present PEDOT:PSS/ZnO hybrid composite sheets were exam­


ined as a sensing medium for wearable thermal sensors, which can be
Fig. 6. Scanning electron microscope (SEM) images for the cross-sectional parts
of the PEDOT:PSS/ZnO hybrid composite sheets with three different sizes of another candidate for the applications of OTEDs. As shown in Fig. 9a,
ZnO particles (DZnO = 0.1, 5, 45 μm). Note that the dashed circles in pink the 500 μm-thick PEDOT:PSS/ZnO hybrid composite sheets (films) (70
indicate the ZnO particles placed in the hybrid composite sheets. (For inter­ wt%, 45 μm ZnO) were employed for the fabrication of the wearable
pretation of the references to colour in this figure legend, the reader is referred thermal sensors on thin copper foils (thickness = 50 μm) with a 50 μm-
to the web version of this article.) thick cellophane tape encapsulation. The basic thermoelectric

6
W. Lee et al. Chemical Engineering Journal 415 (2021) 128935

Fig. 7. Illustration for the role of ZnO particles in the PEDOT:PSS/ZnO hybrid composites with respect to the size of ZnO particles. Yellow circles stand for the ZnO
particles, while black and navy thick lines represent the PEDOT:PSS flakes. The bridging role of ZnO particles between the PEDOT:PSS flakes is expressed by the red
arrows, which might induce the closely contacted PEDOT:PSS flakes as marked with the dashed pink circles. (For interpretation of the references to colour in this
figure legend, the reader is referred to the web version of this article.)

Fig. 8. (a) Work function and (b) Raman spectra for


the PEDOT:PSS/ZnO hybrid composite films as a
function of ZnO content (DZnO = 45 μm). Note that
the main peak at ca. 1430 cm− 1 corresponds to the
symmetric stretching of Cα-Cβ bonds in the thiophene
rings of PEDOT. (c) XPS spectra for the PEDOT:PSS/
ZnO films according to the size of ZnO particles (70
wt%): (left) Zn2p and (right) S2p peaks. Arrows in
pink denote the direction of peak shift. (For inter­
pretation of the references to colour in this figure
legend, the reader is referred to the web version of
this article.)

performances of flexible devices are summarized in Table S2. As shown source. Here it is noted that the sensor did not touch but approached the
in Fig. S9, the thermoelectric characteristics of these flexible devices cool source. Interestingly, the current signal turned to a negative value
were measured upon bending by varying the radius of curvature (RCV). by further keeping the fingertip device near the cool source. This can be
Note that RCV = 5 mm means a bended state smaller than the perimeter explained by the fact that the Cu electrode on the air side was changed to
of an average adult’s little finger. The electrical conductivity (and cool from hot, even though the Cu electrode contacting the finger sur­
output current) was slightly increased at RCV = 5 mm compared to that face might be unchanged due to the constant body temperature. As
before bending (flat), but it became marginally lower by further shown in Fig. 9c, the voltage signals showed the similar trend as
increasing the radius up to RCV = 20 mm (see the top panel in Fig. S9). measured for the current signals. These results support that the OTEDs
However, the Seebeck coefficient (see output voltage) was relatively not with the present PEDOT:PSS/ZnO hybrid composite sheets can be
so much different with the extent of bending (see the middle panel in applicable for artificial skins in humanoid robotics and/or protective
Fig. S9). As a result, a slightly lowered value in power factor was ob­ gears in thermally hazardous working areas [45–47].
tained at the higher radius of curvature (this was more pronounced at
lower ΔT than higher ΔT (see the bottom panel in Fig. S9)). Then, the 4. Conclusions
fabricated flexible devices were conformally attached to the fingertip by
bending nicely. First, the current signal was measured by approaching The PEDOT:PSS/ZnO hybrid composites were successfully prepared
the fingertip with the thermal sensor toward a hot source and then by varying the content and size of ZnO particles. The mixture solutions
moving it to a cool source (see Fig. S10). When the fingertip device of PEDOT:PSS and ZnO became viscous upon stirring and underwent
approached to the hot source (candle), the device current was very gelation after heating for 1 h, which was assigned to the interactions
quickly increased as observed in Fig. 9b. Upon moving the fingertip between hydroxyl groups (and/or Zn2+) in ZnO particles and sulfonic
device to the cool source (ice pack) from the candle side, the current was acid groups (and/or SO−3 ) in PSS chains (as measured by XPS). The
decreased fast and maintained as the device kept onside of the cool bigger ZnO particles delivered higher current (IOUT) from the OTEDs

7
W. Lee et al. Chemical Engineering Journal 415 (2021) 128935

Fig. 9. (a) Flexible thermal sensor (right: cross-


sectional structure), which is attached to the
fingertip, based on the OTEDs with the PEDOT:PSS/
ZnO hybrid composite sheets (DZnO = 45 μm and ZnO
content = 70 wt%). Note that the photograph was
slightly blurred to make the fingerprint unrecogniz­
able. The spacer in the cross-sectional structure was
prepared using a cellophane film. (b) Current and (c)
voltage responses of the flexible thermal sensor upon
approaching and retracting the sensor-attached
fingertip toward hot and cool sources (red arrows:
approaching toward the hot source; black arrows:
retracting from the hot source; blue arrows:
approaching toward the cool source; pink arrows:
retracting from the cool source) (see Fig. S10 and
video clips: Fig_S10_I.mp4 and Fig_S10_V.mp4). (For
interpretation of the references to colour in this
figure legend, the reader is referred to the web
version of this article.)

with the PEDOT:PSS/ZnO composites, while the device voltage (VOUT) References
was affected only by the content of ZnO particles. The resulting power
(POUT) and power factor (PF) of the OTEDs with the PEDOT:PSS/ZnO [1] H. Yang, D. Qi, Z. Liu, B.K. Chandran, T. Wang, J. Yu, X. Chen, Adv. Mater. 28
(2016) 9175–9181.
composites (70 wt% of 45 μm ZnO) reached 0.55 nW and 1.7 × 10− 6 μW [2] F. Zhang, Y. Zang, D. Huang, C.-A. Di, D. Zhu, Nat. Commun. 6 (2015) 8356.
m− 1K− 2 at △T = 50 K, respectively (note that very low power of only [3] H. Cheng, Y. Du, B. Wang, Z. Mao, H. Xu, L. Zhang, Y. Zhong, W. Jiang, L. Wang,
0.07 nW was generated from the OTEDs with the pristine PEDOT:PSS X. Sui, Chem. Eng. J. 338 (2018) 1–7.
[4] Y. Jia, L. Shen, J. Liu, W. Zhou, Y. Du, J. Xu, C. Liu, G. Zhang, Z. Zhang, F. Jiang,
sheets). The microscopic morphology study disclosed that the larger J. Mater. Chem. C 7 (2019) 3496.
ZnO particles could play a better bridging role between the flakes of [5] S. Han, F. Jiao, Z.U. Khan, J. Edberg, S. Fabiano, X. Crispin, Adv. Funct. Mater. 27
PEDOT:PSS/ZnO composites leading to the enhanced electrical con­ (2017) 1703549.
[6] X. Dong, S. Xiong, B. Luo, R. Ge, Z. Li, J. Li, Y. Zhou, A.C.S. Appl, Mater. Interfaces
ductivity in the composite sheets. The Raman spectra evidenced the 10 (2018) 26687–26693.
extended conjugation length in the PEDOT chains by the presence of [7] J.H. We, S.J. Kim, B.J. Cho, Energy 73 (2014) 506–512.
ZnO particles, which contributed to the enhanced electrical conductivity [8] R. Tian, C. Wan, Y. Wang, Q. Wei, T. Ishida, A. Yamamoto, A. Tsuruta, W. Shin,
S. Li, K. Koumoto, J. Mater. Chem. A 5 (2017) 564.
leading to the improved thermoelectric performances for the OTEDs
[9] S.D. Kang, G.J. Snyder, Nat. Mater. 16 (2017) 252–257.
with the PEDOT:PSS/ZnO hybrid composites. The PEDOT:PSS/ZnO [10] Z.U. Khan, J. Edberg, M.M. Hamedi, R. Gabrielsson, H. Granberg, L. Wågberg,
composite sheets could be successfully applied as a flexible/wearable I. Engquist, M. Berggren, X. Crispin, Adv. Mater 28 (2016) 4556–4562.
thermal sensor, which can be attachable to a fingertip, with an instant [11] W. Glatz, S. Muntwyler, C. Hierold, Sens. Actuator. A 132 (2006) 337–345.
[12] Y. Wang, J. Yang, L. Wang, K. Du, Q. Yin, Q. Yin, A.C.S. Appl, Mater. Interfaces 9
and fast thermal response toward hot and cool sources. Therefore, it is (2017) 20124–20131.
expected that the present PEDOT:PSS/ZnO composite sheets can be used [13] H. Li, S. Liu, P. Li, D. Yuan, X. Zhou, J. Sun, X. Lu, C. He, Carbon 136 (2018)
as one of the useful thermoelectric materials for various types of ther­ 292–298.
[14] Y.H. Kim, C. Sachse, M.L. Machala, C. May, L. Müller-Meskamp, K. Leo, Adv. Funct.
moelectric devices. Mater. 21 (2011) 1076–1081.
[15] D. Alemu, H.-Y. Wei, K.-C. Ho, C.-W. Chu, Energy Environ. Sci. 5 (2012) 9662.
[16] Z. Yu, Y. Xia, D. Du, J. Ouyang, ACS Appl. Mater. Interfaces 8 (2016)
Declaration of Competing Interest 11629–11638.
[17] S. Xu, M. Hong, X.-L. Shi, Y. Wang, L. Ge, Y. Bai, L. Wang, M. Dargusch, J. Zou, Z.-
The authors declare that they have no known competing financial G. Chen, Chem. Mater. 31 (2019) 5238–5244.
[18] Z. Fan, P. Li, D. Du, J. Ouyang, Adv. Energy Mater. 7 (2017) 1602116.
interests or personal relationships that could have appeared to influence [19] E.J. Bae, Y.H. Kang, K.-S. Jang, S.Y. Cho, Sci. Rep. 6 (2016) 18805.
the work reported in this paper. [20] S. Liu, H. Li, C. He, Carbon 149 (2019) 25–32.
[21] C.S. Kim, H.M. Yang, J. Lee, G.S. Lee, H. Choi, Y.J. Kim, S.H. Lim, S.H. Cho, B.
J. Cho, ACS Energy Lett. 3 (2018) 501–507.
Acknowledgements [22] F.J. Lesage, R. Pelletier, L. Fournier, É.V. Sempels, Energy Convers. Manage. 74
(2013) 51–59.
This work was financially supported by the National Research [23] X.F. Zheng, Y.Y. Yan, K. Simpson, Appl. Therm. Eng. 53 (2013) 305–311.
[24] S.A. Whalen, R.C. Dykhuizen, Energy Convers. Manage. 64 (2012) 397–402.
Foundation (NRF) of Korea (2018R1D1A3B07046214, [25] Y. Yang, X.-J. Wei, J. Liu, J. Phys. D Appl. Phys. 40 (2007) 5790–5800.
2018R1D1A1B07051075, Basic Science Research Pro­ [26] B.S. Tomar, A. Shahin, M.S. Tirumkudulu, Soft Matter 16 (2020) 3476–3484.
gram_2018R1A6A1A03024962) and the Ministry of Trade, Industry and [27] S.S. Shojaie, W.B. Krantz, A.R. Greenberg, J. Membrane Sci. 94 (1994) 281–298.
[28] H. Yi, D. Wang, L. Duan, F. Haque, C. Xu, Y. Zhang, G. Conibeer, A. Uddin,
Energy (MOTIE) - Korea Institute for Advancement of Technology
Electrochim. Acta 319 (2019) 349–358.
(KIAT) through the International Cooperative R&D program (Project [29] J. Zhou, D.H. Anjum, L. Chen, X. Xu, I.A. Ventura, L. Jiang, G. Lubineau, J. Mater.
No. P0011262). Chem. C 2 (2014) 9903–9910.
[30] T.-C. Tsai, H.-C. Chang, C.-H. Chen, W.-T. Whang, Org. Electron. 12 (2011)
2159–2164.
Appendix A. Supplementary data [31] K. Kirihara, Q. Wei, M. Mukaida, T. Ishida, Synth. Met. 225 (2017) 41–48.
[32] M.P. Gordon, E.W. Zaia, P. Zhou, B. Russ, N.E. Coates, A. Sahu, J.J. Urban, J. Appl.
Supplementary data to this article can be found online at https://doi. Polym. Sci. 134 (2017) 44070.
[33] F. Batool, V. Bindiganavile, Constr. Build. Mater. 149 (2017) 17–28.
org/10.1016/j.cej.2021.128935.

8
W. Lee et al. Chemical Engineering Journal 415 (2021) 128935

[34] Z. Wang, T. Iizuka, M. Kozako, Y. Ohki, T. Tanaka, IEEE Trans. Dielectr. Electr. [41] C.-H. Lin, C.-W. Huang, P.-H. Wang, T.-F. Guo, T.-C. Wen, J. Taiwan Inst. Chem.
Insul. 18 (2011) 1963–1972. Eng. 107 (2020) 72–78.
[35] A. Tabib, N. Sdiri, H. Elhouichet, M. Férid, J. Alloys Compd. 622 (2015) 687–694. [42] B. Dilasari, Y. Jung, K. Kwon, J. Ind. Eng. Chem. 45 (2017) 375–379.
[36] M. Caglar, S. Ilican, Y. Caglar, F. Yakuphanoglu, Appl. Surf. Sci. 255 (2009) [43] H. Park, S.H. Lee, F.S. Kim, H.H. Choi, I.W. Cheong, J.H. Kim, J. Mater. Chem. A 2
4491–4496. (2014) 6532–6539.
[37] P. Sathishkumar, Z. Li, R. Govindan, R. Jayakumar, C. Wang, F.L. Gu, Appl. Surf. [44] W. Lee, M. Song, S. Park, S. Nam, J. Seo, H. Kim, Y. Kim, Sci. Rep. 6 (2016) 33795.
Sci. 536 (2021), 147741. [45] V. Leonov, IEEE Sens. J. 13 (2013) 2284–2291.
[38] M.M. Rahman, Polymers 12 (2020) 1535. [46] J. Yuan, R. Zhu, G. Li, Adv. Mater. Technol. 5 (2020) 2000419.
[39] I. Koresh, Y. Amouyal, J. Eur. Ceram. Soc. 37 (2017) 3541–3550. [47] P. Zhu, Y. Wang, Y. Wang, H. Mao, Q. Zhang, Y. Deng, Adv. Energy Mater. 10
[40] P. Jood, R.J. Mehta, Y. Zhang, G. Peleckis, X. Wang, R.W. Siegel, T. Borca-Tasciuc, (2020) 2001945.
S.X. Dou, G. Ramanath, Nano Lett. 11 (2011) 4337–4342.

You might also like